This page intentionally left blank irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page i accessible, affordable, active learning :LOH\3/86LVDQLQQRYDWLYHUHVHDUFKEDVHGRQOLQHHQYLURQPHQWIRU HIIHFWLYHWHDFKLQJDQGOHDUQLQJ :LOH\3/86 «PRWLYDWHVVWXGHQWVZLWK FRQÀGHQFHERRVWLQJ IHHGEDFNDQGSURRIRI SURJUHVV «VXSSRUWVLQVWUXFWRUVZLWK UHOLDEOHUHVRXUFHVWKDW UHLQIRUFHFRXUVHJRDOV LQVLGHDQGRXWVLGHRI WKHFODVVURRP ,QFOXGHV ,QWHUDFWLYH 7H[WERRN 5HVRXUFHV :LOH\3/86/HDUQ0RUH ZZZZLOH\SOXVFRP irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page ii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irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page iii BASIC ENGINEERING CIRCUIT ANALYSIS This page intentionally left blank irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page v BASIC ENGINEERING CIRCUIT ANALYSIS Tenth Edition J. DAVID IRWIN Auburn University R. MARK NELMS Auburn University John Wiley & Sons, Inc. irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page vi Vice President and Executive Publisher Associate Publisher Production Manager Senior Production Editor Marketing Manager Creative Director Senior Designer Production Management Services Senior Illustration Editor Senior Photo Editor Editorial Assistant Executive Media Editor Cover Photo Don Fowley Dan Sayre Dorothy Sinclair Valerie A. Vargas Christopher Ruel Harry Nolan Kevin Murphy Furino Production Anna Melhorn Lisa Gee Katie Singleton Thomas Kulesa provided courtesy of Tesla Motors, Inc. This book was set in 10/12 Times by Prepare and printed and bound by Courier-Kendallville. The cover was printed by Courier-Kendallville. Copyright © 2011, 2008, 2005 John Wiley & Sons, Inc. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise, except as permitted under Sections 107 or 108 of the 1976 United States Copyright Act, without either the prior written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to the Copyright Clearance Center, Inc. 222 Rosewood Drive, Danvers, MA 01923, website www.copyright.com. Requests to the Publisher for permission should be addressed to the Permissions Department, John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030-5774, (201)748-6011, fax (201)748-6008, website http://www.wiley.com/go/permissions. Founded in 1807, John Wiley & Sons, Inc. has been a valued source of knowledge and understanding for more than 200 years, helping people around the world meet their needs and fulfill their aspirations. Our company is built on a foundation of principles that include responsibility to the communities we serve and where we live and work. In 2008, we launched a Corporate Citizenship Initiative, a global effort to address the environmental, social, economic, and ethical challenges we face in our business. Among the issues we are addressing are carbon impact, paper specifications and procurement, ethical conduct within our business and among our vendors, and community and charitable support. For more information, please visit our website: www.wiley.com/go/citizenship. Evaluation copies are provided to qualified academics and professionals for review purposes only, for use in their courses during the next academic year. These copies are licensed and may not be sold or transferred to a third party. Upon completion of the review period, please return the evaluation copy to Wiley. Return instructions and a free of charge return shipping label are available at www.wiley.com/go/returnlabel. Outside of the United States, please contact your local representative. ISBN-13 978-0-470-63322-9 Printed in the United States of America 10 9 8 7 6 5 4 3 2 1 irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page vii To my loving family: Edie Geri, Bruno, Andrew and Ryan John, Julie, John David and Abi Laura To my parents: Robert and Elizabeth Nelms This page intentionally left blank irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page ix BRIEF CONTENTS 1 CHAPTER 2 CHAPTER 3 CHAPTER 4 CHAPTER 5 CHAPTER 6 CHAPTER 7 CHAPTER 8 CHAPTER 9 C H A P T E R 10 C H A P T E R 11 C H A P T E R 12 C H A P T E R 13 C H A P T E R 14 CHAPTER 15 C H A P T E R 16 CHAPTER Basic Concepts Resistive Circuits 1 25 Nodal and Loop Analysis Techniques 102 Operational Amplifiers 156 Additional Analysis Techniques 189 Capacitance and Inductance 245 First- and Second-Order Transient Circuits 296 AC Steady-State Analysis 369 Steady-State Power Analysis 435 Magnetically Coupled Networks 491 Polyphase Circuits 541 Variable-Frequency Network Performance 577 The Laplace Transform 667 Application of the Laplace Transform to Circuit Analysis 695 Fourier Analysis Techniques 751 Two-Port Networks 801 This page intentionally left blank irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page xi CONTENTS Preface CHAPTER CHAPTER 1 System of Units 2 Basic Quantities 2 Circuit Elements 8 Summary 17 Problems 17 3.1 3.2 3.3 3.4 2 RESISTIVE CIRCUITS 2.1 2.2 2.3 2.4 2.5 2.6 Ohm’s Law 26 Kirchhoff’s Laws 31 Single-Loop Circuits 39 Single-Node-Pair Circuits 46 Series and Parallel Resistor Combinations 51 Circuits with Series-Parallel Combinations of Resistors 55 2.7 Wye Delta Transformations 61 2.8 Circuits with Dependent Sources 64 2.9 Resistor Technologies for Electronic Manufacturing 69 2.10 Application Examples 72 2.11 Design Examples 75 Summary 81 Problems 82 4 OPERATIONAL AMPLIFIERS 25 4.1 4.2 4.3 4.4 4.5 4.6 156 Introduction 157 Op-Amp Models 157 Fundamental Op-Amp Circuits 163 Comparators 172 Application Examples 173 Design Examples 176 Summary 179 Problems 180 CHAPTER 5 ADDITIONAL ANALYSIS TECHNIQUES 5.1 5.2 5.3 102 Nodal Analysis 102 Loop Analysis 122 Application Example 137 Design Example 139 Summary 139 Problems 140 CHAPTER CHAPTER 3 NODAL AND LOOP ANALYSIS TECHNIQUES 1 BASIC CONCEPTS 1.1 1.2 1.3 xiii Introduction 190 Superposition 192 Thévenin’s and Norton’s Theorems 198 189 irwin_fm_i-xxii-hr.qxd xii 5.4 5.5 5.6 3-09-2010 15:50 Page xii CONTENTS Maximum Power Transfer 216 Application Example 220 Design Examples 221 Summary 227 Problems 227 8.8 Analysis Techniques 396 8.9 Application Examples 408 8.10 Design Examples 410 Summary 413 Problems 414 CHAPTER CHAPTER 6 CAPACITANCE AND INDUCTANCE 6.1 6.2 6.3 6.4 6.5 6.6 Sinusoids 370 Sinusoidal and Complex Forcing Functions 373 Phasors 377 Phasor Relationships for Circuit Elements 379 Impedance and Admittance 383 Phasor Diagrams 390 Basic Analysis Using Kirchhoff’s Laws 393 10 MAGNETICALLY COUPLED NETWORKS 10.1 10.2 10.3 10.4 10.5 10.6 369 491 Mutual Inductance 492 Energy Analysis 503 The Ideal Transformer 506 Safety Considerations 515 Application Examples 516 Design Examples 521 Summary 525 Problems 526 CHAPTER 11 POLYPHASE CIRCUITS 11.1 11.2 11.3 11.4 11.5 435 Instantaneous Power 436 Average Power 437 Maximum Average Power Transfer 442 Effective or rms Values 447 The Power Factor 450 Complex Power 452 Power Factor Correction 457 Single-Phase Three-Wire Circuits 461 Safety Considerations 464 Application Examples 472 Design Examples 476 Summary 478 Problems 478 CHAPTER 8 AC STEADY-STATE ANALYSIS 8.1 8.2 8.3 8.4 8.5 8.6 8.7 296 Introduction 297 First-Order Circuits 298 Second-Order Circuits 319 Application Examples 333 Design Examples 343 Summary 351 Problems 351 CHAPTER 9.1 9.2 9.3 9.4 9.5 9.6 9.7 9.8 9.9 9.10 9.11 7 FIRST- AND SECOND-ORDER TRANSIENT CIRCUITS 7.1 7.2 7.3 7.4 7.5 245 Capacitors 246 Inductors 254 Capacitor and Inductor Combinations 264 RC Operational Amplifier Circuits 272 Application Examples 274 Design Examples 279 Summary 280 Problems 281 CHAPTER 9 STEADY-STATE POWER ANALYSIS Three-Phase Circuits 542 Three-Phase Connections 547 Source/Load Connections 548 Power Relationships 557 Power Factor Correction 561 541 irwin_fm_i-xxii-hr2.qxd 6-09-2010 14:10 Page xiii CONTENTS 11.6 Application Examples 562 11.7 Design Examples 566 Summary 570 Problems 570 CHAPTER 12 VARIABLE-FREQUENCY NETWORK PERFORMANCE 12.1 12.2 12.3 12.4 12.5 12.6 12.7 CHAPTER 15.1 15.2 15.3 15.4 667 14 APPLICATION OF THE LAPLACE TRANSFORM TO CIRCUIT ANALYSIS 16 TWO-PORT NETWORKS 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8 801 Admittance Parameters 802 Impedance Parameters 805 Hybrid Parameters 807 Transmission Parameters 809 Parameter Conversions 811 Interconnection of Two-Ports 811 Application Examples 815 Design Example 819 Summary 821 Problems 821 APPENDIX 695 COMPLEX NUMBERS Index 14.1 Laplace Circuit Solutions 696 14.2 Circuit Element Models 697 751 Fourier Series 752 Fourier Transform 773 Application Examples 780 Design Example 787 Summary 793 Problems 794 CHAPTER Definition 668 Two Important Singularity Functions 669 Transform Pairs 671 Properties of the Transform 673 Performing the Inverse Transform 676 Convolution Integral 681 Initial-Value and Final-Value Theorems 685 Application Examples 687 Summary 689 Problems 689 CHAPTER 15 FOURIER ANALYSIS TECHNIQUES 13 THE LAPLACE TRANSFORM Analysis Techniques 699 Transfer Function 712 Pole-Zero Plot/Bode Plot Connection 724 Steady-State Response 727 Application Examples 729 Design Examples 731 Summary 738 Problems 738 577 Variable Frequency-Response Analysis 578 Sinusoidal Frequency Analysis 586 Resonant Circuits 597 Scaling 619 Filter Networks 620 Application Examples 645 Design Examples 649 Summary 655 Problems 656 CHAPTER 13.1 13.2 13.3 13.4 13.5 13.6 13.7 13.8 14.3 14.4 14.5 14.6 14.7 14.8 xiii 829 834 This page intentionally left blank irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page xv PREFACE Circuit analysis is not only fundamental to the entire breadth of electrical and computer engineering—the concepts studied here extend far beyond those boundaries. For this reason it remains the starting point for many future engineers who wish to work in this field. The text and all the supplementary materials associated with it will aid you in reaching this goal. We strongly recommend while you are here to read the Preface closely and view all the resources available to you as a learner. And one last piece of advice, learning requires practice and repetition, take every opportunity to work one more problem or study one more hour than you planned. In the end, you’ll be thankful you did. To the Student The Tenth Edition has been prepared based on a careful examination of feedback received from instructors and students. The revisions and changes made should appeal to a wide variety of instructors. We are aware of significant changes taking place in the way this material is being taught and learned. Consequently, the authors and the publisher have created a formidable array of traditional and non-traditional learning resources to meet the needs of students and teachers of modern circuit analysis. To the Instructor • • • • A four-color design is employed to enhance and clarify both text and illustrations. This sharply improves the pedagogical presentation, particularly with complex illustrations. For example, see Figure 2.5 on page 31. New chapter previews provide motivation for studying the material in the chapter. See page 25 for a chapter preview sample. Learning objectives for each chapter have been updated and appear as part of the new chapter openers. End of chapter homework problems have been substantially revised and augmented. There are now approximately 1400 problems in the Tenth Edition, of which over 400 are new! Multiple-choice Fundamentals of Engineering (FE) Exam problems also appear at the end of each chapter. Practical applications have been added for nearly every topic in the text. Since these are items students will naturally encounter on a regular basis, they serve to answer questions such as, “Why is this important?” or “How am I going to use what I learn from this course?” For a typical example application, see page 333. Highlights of the Tenth Edition irwin_fm_i-xxii-hr.qxd xvi 3-09-2010 15:50 Page xvi P R E FAC E • • • • Problem Solving videos have been created showing students step-by-step how to solve all Learning Assessment problems within each chapter. This is a special feature that should significantly enhance the learning experience for each subsection in a chapter. The problem-solving videos (PSVs) are now also available for the Apple iPod. In order to provide maximum flexibility, online supplements contain solutions to examples in the book using MATLAB, PSPICE or MultiSim. The worked examples can be supplied to students as digital files, or one or more of them can be incorporated into custom print editions of the text, depending upon the instructor’s preference. Problem-Solving Strategies have been retained in the Tenth Edition. They are utilized as a guide for the solutions contained in the PSVs. The WileyPLUS resources have been greatly updated and expanded, with additional algorithmic problems, problem-solving videos and much more. New Reading Quiz questions give instructors the opportunity to track student reading and measure their comprehension. New Math Skills Assessments provide faculty with tools to assess students’ mastery of essential mathematical concepts. Not only can faculty measure their students’ math comprehension at the beginning of the term, they also now have resources to which they can direct students to help them reinforce areas where they need to upgrade their skills. Organization This text is suitable for a one-semester, a two-semester or a three-quarter course sequence. The first seven chapters are concerned with the analysis of dc circuits. An introduction to operational amplifiers is presented in Chapter 4. This chapter may be omitted without any loss of continuity; a few examples and homework problems in later chapters must be skipped. Chapters 8–12 are focused on the analysis of ac circuits beginning with the analysis of singlefrequency circuits (single-phase and three-phase) and ending with variable-frequency circuit operation. Calculation of power in single-phase and three-phase ac circuits is also presented. The important topics of the Laplace transform, Fourier transform, and two-port networks are covered in Chapters 13–16. The organization of the text provides instructors maximum flexibility in designing their courses. One instructor may choose to cover the first seven chapters in a single semester, while another may omit Chapter 4 and cover Chapters 1–3 and 5–8. Other instructors have chosen to cover Chapters 1–3, 5–6, and sections 7.1 and 7.2 and then cover Chapters 8 and 9. The remaining chapters can be covered in a second semester course. Text Pedagogy The pedagogy of this text is rich and varied. It includes print and media and much thought has been put into integrating its use. To gain the most from this pedagogy, please review the following elements commonly available in most chapters of this book. Learning Objectives are provided at the outset of each chapter. This tabular list tells the reader what is important and what will be gained from studying the material in the chapter. Examples are the mainstay of any circuit analysis text and numerous examples have always been a trademark of this textbook. These examples provide a more graduated level of presentation with simple, medium and challenging examples. Besides regular examples, numerous Design Examples and Application Examples are found throughout the text. See for example, page 343. Hints can often be found in the page margins. They facilitate understanding and serve as reminders of key issues. See for example, page 6. irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page xvii P R E FAC E xvii Learning Assessments are a critical learning tool in this text. These exercises test the cumulative concepts to that point in a given section or sections. Not only is the answer provided, but a problem-solving video accompanies each of these exercises, demonstrating the solution in step-by-step detail. The student who masters these is ready to move forward. See for example, page 7. Problem-Solving Strategies are step-by-step problem-solving techniques that many students find particularly useful. They answer the frequently asked question, “where do I begin?” Nearly every chapter has one or more of these strategies, which are a kind of summation on problem-solving for concepts presented. See for example, page 121. The Problems have been greatly revised for the 10th Edition. This edition has over 400 new problems of varying depth and level. Any instructor will find numerous problems appropriate for any level class. There are approximately 1400 problems in the 10th Edition! Included with the Problems are FE Exam Problems for each chapter. If you plan on taking the FE Exam, these problems closely match problems you will typically find on the FE Exam. Circuit Simulation and Analysis Software represents a fundamental part of engineering circuit design today. Software such as PSPICE®, MultiSim® and MATLAB® allow engineers to design and simulate circuits quickly and efficiently. As an enhancement with enormous flexibility, all three of these software packages can be employed in the 10th edition. In each case, online supplements are available that contain the solutions to numerous examples in each of these software programs. Instructors can opt to make this material available online or as part of a customized print edition, making this software an integral and effective part of the presentation of course material. The rich collection of material that is provided for this edition offers a distinctive and helpful way for exploring the book’s examples and exercises from a variety of simulation techniques. WileyPLUS is an innovative, research-based, online environment for effective teaching and learning. W H AT D O S T U D E N T S R EC E I V E W I T H W I L E Y P LU S ? A Research-based Design. WileyPLUS provides an online environment that integrates relevant resources, including the entire digital textbook, in an easy-to-navigate framework that helps students study more effectively. • • • WileyPLUS adds structure by organizing textbook content into smaller, more manageable “chunks”. Related media, examples, and sample practice items reinforce the learning objectives. Innovative features such as calendars, visual progress tracking and self-evaluation tools improve time management and strengthen areas of weakness. One-on-one Engagement. With WileyPLUS, students receive 24/7 access to resources that promote positive learning outcomes. Students engage with related examples (in various media) and sample practice items, including: • • • • • FE Exam Questions Reading Quiz Questions Circuit Solutions Learning Assessments Math Skills Assessments WileyPLUS irwin_fm_i-xxii-hr.qxd xviii 3-09-2010 15:50 Page xviii P R E FAC E Measurable Outcomes. Throughout each study session, students can assess their progress and gain immediate feedback. WileyPLUS provides precise reporting of strengths and weaknesses, as well as individualized quizzes, so that students are confident they are spending their time on the right things. With WileyPLUS, students always know the exact outcome of their efforts. W H AT D O I N S T R U C TO R S R EC E I V E W I T H W I L E Y P LU S ? WileyPLUS provides reliable, customizable resources that reinforce course goals inside and outside of the classroom as well as visibility into individual student progress. Pre-created materials and activities help instructors optimize their time. Customizable Course Plan: WileyPLUS comes with a pre-created Course Plan designed by a subject matter expert uniquely for this course. Simple drag-and-drop tools make it easy to assign the course plan as-is or modify it to reflect your course syllabus. Pre-created Activity Types include: • • • • • Questions Readings and Resources Presentation Print Tests Concept Mastery Course Materials and Assessment Content: • • • • • • Lecture Notes PowerPoint Slides Image Gallery Instructor’s Manual Gradable Reading Assignment Questions (embedded with online text) Question Assignments: all end-of-chapter problems coded algorithmically with hints, links to text, whiteboard/show work feature and instructor controlled problem solving help. Gradebook: WileyPLUS provides instant access to reports on trends in class performance, student use of course materials and progress towards learning objectives, helping inform decisions and drive classroom discussions. WileyPLUS. Learn more at www.wileyplus.com. Powered by proven technology and built on a foundation of cognitive research, WileyPLUS has enriched the education of millions of students, in over 20 countries around the world. Supplements The supplements list is extensive and provides instructors and students with a wealth of traditional and modern resources to match different learning needs. Problem-Solving Videos are offered again in the 10th Edition in an iPod-compatible format. The videos provide step-by-step solutions to Learning Assessments. Videos for Learning Assessments will follow directly after a chapter feature called Problem-Solving Strategy. Students who have used these videos with past editions have found them to be very helpful. The Solutions Manual for the 10th Edition has been completely redone, checked and double-checked for accuracy. Although it is hand-written to avoid typesetting errors, it is the most accurate solutions manual ever created for this textbook. Qualified instructors who adopt the text for classroom use can download it off Wiley’s Instructor’s Companion Site. irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page xix P R E FAC E xix PowerPoint Lecture Slides are an especially valuable supplementary aid for some instructors. While most publishers make only figures available, these slides are true lecture tools that summarize the key learning points for each chapter and are easily editable in PowerPoint. The slides are available for download from Wiley’s Instructor Companion Site for qualified adopters. Over the more than two decades that this text has been in existence, we estimate more than one thousand instructors have used our book in teaching circuit analysis to hundreds of thousand of students. As authors there is no greater reward than having your work used by so many. We are grateful for the confidence shown in our text and for the numerous evaluations and suggestions from professors and their students over the years. This feedback has helped us continuously improve the presentation. For this Tenth edition, we especially thank Jim Rowland from the University of Kansas for his assistance with the chapter openers and Stephen Haddock with Auburn University for his assistance with PSPICE®, MultiSim® and MATLAB® supplemental materials. The authors also wish to express a special thanks to Sandy Johnson for her diligence and dedication in the preparation of this 10th edition. We were fortunate to have an outstanding group of faculty who has participated in reviews, surveys and focus groups for this edition. They are: Jorge Aravena, Louisiana State University James Conrad, University of North Carolina, Charlotte Paul King, Vanderbilt University Gordon Lee, San Diego State University Tokunbo Ogunfunmi, Santa Clara University Michael Polis, Oakland University The preparation of this book and the materials that support it have been handled with both enthusiasm and great care. The combined wisdom and leadership of our colleagues at Wiley has resulted in a tremendous team effort that has addressed every aspect of the presentation. This team included the following individuals: Executive Publisher, Don Fowley Associate Publisher, Dan Sayre Executive Media Editor, Tom Kulesa Executive Marketing Manager, Chris Ruel Senior Production Editor, Valerie Vargas Senior Designer, Kevin Murphy Production Manager, Dorothy Sinclair Senior Photo Editor, Lisa Gee Media Editor, Lauren Sapira Editorial Assistant, Katie Singelton Each member of this team played a vital role in preparing the package that is the Tenth Edition of Basic Engineering Circuit Analysis. We are most appreciative of their many contributions. As in the past, we are most pleased to acknowledge the support that has been provided by numerous individuals to earlier editions of this book. Our Auburn colleagues who have helped are: Thomas A. Baginski Travis Blalock Henry Cobb Acknowledgments irwin_fm_i-xxii-hr.qxd xx 3-09-2010 15:50 Page xx P R E FAC E Bill Dillard Zhi Ding Kevin Driscoll E. R. Graf L. L. Grigsby Charles A. Gross Stephen Haddock David C. Hill M. A. Honnell R. C. Jaeger Keith Jones Betty Kelley Ray Kirby Matthew Langford Aleck Leedy George Lindsey Jo Ann Loden James L. Lowry David Mack Paulo R. Marino M. S. Morse Sung-Won Park John Parr Monty Rickles C. L. Rogers Tom Shumpert Les Simonton James Trivltayakhum Susan Williamson Jacinda Woodward Many of our friends throughout the United States, some of whom are now retired, have also made numerous suggestions for improving the book: David Anderson, University of Iowa Jorge Aravena, Louisiana State University Les Axelrod, Illinois Institute of Technology Richard Baker, UCLA Charles F. Bunting, Oklahoma State University John Choma, University of Southern California David Conner, University of Alabama at Birmingham James L. Dodd, Mississippi State University Kevin Donahue, University of Kentucky John Durkin, University of Akron Prasad Enjeti, Texas A&M University Earl D. Eyman, University of Iowa Arvin Grabel, Northeastern University Paul Gray, University of Wisconsin-Platteville Ashok Goel, Michigan Technological University Walter Green, University of Tennessee Paul Greiling, UCLA Mohammad Habli, University of New Orleans John Hadjilogiou, Florida Institute of Technology Yasser Hegazy, University of Waterloo Keith Holbert, Arizona State University irwin_fm_i-xxii-hr.qxd 3-09-2010 15:50 Page xxi P R E FAC E Aileen Honka, The MOSIS Service- USC Inf. Sciences Institute Marty Kaliski, Cal Poly, San Luis Obispo Muhammad A. Khaliq, Minnesota State University Ralph Kinney, LSU Robert Krueger, University of Wisconsin K. S. P. Kumar, University of Minnesota Jung Young Lee, UC Berkeley student Aleck Leedy, Murray State University Hongbin Li, Stevens Institute of Technology James Luster, Snow College Erik Luther, National Instruments Ian McCausland, University of Toronto Arthur C. Moeller, Marquette University Darryl Morrell, Arizona State University M. Paul Murray, Mississippi State University Burks Oakley II, University of Illinois at Champaign-Urbana John O’Malley, University of Florida Arnost Neugroschel, University of Florida William R. Parkhurst, Wichita State University Peyton Peebles, University of Florida Jian Peng, Southeast Missouri State University Clifford Pollock, Cornell University George Prans, Manhattan College Mark Rabalais, Louisiana State University Tom Robbins, National Instruments Armando Rodriguez, Arizona State University James Rowland, University of Kansas Robert N. Sackett, Normandale Community College Richard Sanford, Clarkson University Peddapullaiah Sannuti, Rutgers University Ronald Schulz, Cleveland State University M. E. Shafeei, Penn State University at Harrisburg Martha Sloan, Michigan Technological University Scott F. Smith, Boise State University Karen M. St. Germaine, University of Nebraska Janusz Strazyk, Ohio University Gene Stuffle, Idaho State University Thomas M. Sullivan, Carnegie Mellon University Saad Tabet, Florida State University Val Tareski, North Dakota State University Thomas Thomas, University of South Alabama Leonard J. Tung, Florida A&M University/Florida State University Marian Tzolov, Lock Haven University Darrell Vines, Texas Tech University Carl Wells, Washington State University Seth Wolpert, University of Maine Finally, Dave Irwin wishes to express his deep appreciation to his wife, Edie, who has been most supportive of our efforts in this book. Mark Nelms would like to thank his parents, Robert and Elizabeth, for their support and encouragement. J. David Irwin and R. Mark Nelms xxi This page intentionally left blank irwin01_001-024hr.qxd 30-06-2010 13:16 Page 1 C H A PT E R BASIC CONCEPTS 1 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Review the SI system of units and standard prefixes ■ Know the definitions of basic electrical quantities: voltage, current, and power ■ Know the symbols for and definitions of independent and dependent sources ■ Be able to calculate the power absorbed by a circuit element using the passive sign convention Courtesy NASA, 2009 H Hubble Space Telescope If you were asked to identify the Dynamics Observatory was launched to aid in studying our top engineering achievements that depend on currents, volt- sun’s dynamic processes including high resolution measure- ages, and power in electrical systems, would NASA’s Hubble ments of solar flares; it is the first mission of NASA’s Living Space Telescope make your list? It should. Launched over 20 with a Star program. years ago into an orbit 375 miles above the Earth’s surface, Sophisticated as it is, the power of the Hubble Space the Hubble Telescope avoids distorting effects of the atmos- Telescope is rooted in the fundamental concepts you will begin phere and gives significant new data about the universe. It to study in this chapter—charge, current, voltage, power, and features multiple channels having many intricate electrical batteries. These core principles are the fundamental building systems that detect different wavelengths of light and blocks of your understanding of electrical engineering and your enables us to examine our solar system as well as remote ability to analyze and design more complicated electrical sys- galaxies. The success of the Hubble Space Telescope program tems. Just as the Hubble has led to even greater innovations, has led to other NASA plans. In February 2010, the Solar we cannot imagine today what else may lie ahead for you. 1 irwin01_001-024hr.qxd 2 30-06-2010 CHAPTER 1 13:16 Page 2 BASIC CONCEPTS 1.1 System of Units Figure 1.1 > The system of units we employ is the international system of units, the Système International des Unités, which is normally referred to as the SI standard system. This system, which is composed of the basic units meter (m), kilogram (kg), second (s), ampere (A), kelvin (K), and candela (cd), is defined in all modern physics texts and therefore will not be defined here. However, we will discuss the units in some detail as we encounter them in our subsequent analyses. The standard prefixes that are employed in SI are shown in Fig. 1.1. Note the decimal relationship between these prefixes. These standard prefixes are employed throughout our study of electric circuits. Circuit technology has changed drastically over the years. For example, in the early 1960s the space on a circuit board occupied by the base of a single vacuum tube was about the size of a quarter (25-cent coin). Today that same space could be occupied by an Intel Pentium integrated circuit chip containing 50 million transistors. These chips are the engine for a host of electronic equipment. 10–12 10–9 10–6 10–3 1 103 106 109 1012 Standard SI prefixes. pico (p) nano (n) micro () milli (m) 1.2 Basic Quantities I1=2 A Before we begin our analysis of electric circuits, we must define terms that we will employ. However, in this chapter and throughout the book our definitions and explanations will be as simple as possible to foster an understanding of the use of the material. No attempt will be made to give complete definitions of many of the quantities because such definitions are not only unnecessary at this level but are often confusing. Although most of us have an intuitive concept of what is meant by a circuit, we will simply refer to an electric circuit as an interconnection of electrical components, each of which we will describe with a mathematical model. The most elementary quantity in an analysis of electric circuits is the electric charge. Our interest in electric charge is centered around its motion, since charge in motion results in an energy transfer. Of particular interest to us are those situations in which the motion is confined to a definite closed path. An electric circuit is essentially a pipeline that facilitates the transfer of charge from one point to another. The time rate of change of charge constitutes an electric current. Mathematically, the relationship is expressed as Circuit 1 i(t) = (a) I2=–3 A Circuit 2 (b) Figure 1.2 Conventional current flow: (a) positive current flow; (b) negative current flow. kilo (k) mega (M) giga (G) tera (T) dq(t) dt t or q(t) = 3-q i(x) dx 1.1 where i and q represent current and charge, respectively (lowercase letters represent time dependency, and capital letters are reserved for constant quantities). The basic unit of current is the ampere (A), and 1 ampere is 1 coulomb per second. Although we know that current flow in metallic conductors results from electron motion, the conventional current flow, which is universally adopted, represents the movement of positive charges. It is important that the reader think of current flow as the movement of positive charge regardless of the physical phenomena that take place. The symbolism that will be used to represent current flow is shown in Fig. 1.2. I1 = 2 A in Fig. 1.2a indicates that at any point in the wire shown, 2 C of charge pass from left to right each second. I2 = -3 A in Fig. 1.2b indicates that at any point in the wire shown, 3 C of charge pass from right to left each second. Therefore, it is important to specify not only the magnitude of the variable representing the current but also its direction. irwin01_001-024hr.qxd 30-06-2010 13:16 Page 3 SECTION 1.2 i(t) BASIC QUANTITIES Figure 1.3 i(t) Two common types of current: (a) alternating current (ac); (b) direct current (dc). t (a) t (b) The two types of current that we encounter often in our daily lives, alternating current (ac) and direct current (dc), are shown as a function of time in Fig. 1.3. Alternating current is the common current found in every household and is used to run the refrigerator, stove, washing machine, and so on. Batteries, which are used in automobiles and flashlights, are one source of direct current. In addition to these two types of currents, which have a wide variety of uses, we can generate many other types of currents. We will examine some of these other types later in the book. In the meantime, it is interesting to note that the magnitude of currents in elements familiar to us ranges from soup to nuts, as shown in Fig. 1.4. We have indicated that charges in motion yield an energy transfer. Now we define the voltage (also called the electromotive force, or potential) between two points in a circuit as the difference in energy level of a unit charge located at each of the two points. Voltage is very similar to a gravitational force. Think about a bowling ball being dropped from a ladder into a tank of water. As soon as the ball is released, the force of gravity pulls it toward the bottom of the tank. The potential energy of the bowling ball decreases as it approaches the bottom. The gravitational force is pushing the bowling ball through the water. Think of the bowling ball as a charge and the voltage as the force pushing the charge through a circuit. Charges in motion represent a current, so the motion of the bowling ball could be thought of as a current. The water in the tank will resist the motion of the bowling ball. The motion of charges in an electric circuit will be impeded or resisted as well. We will introduce the concept of resistance in Chapter 2 to describe this effect. Work or energy, w(t) or W, is measured in joules (J); 1 joule is 1 newton meter (N ⭈ m). Hence, voltage [v(t) or V] is measured in volts (V) and 1 volt is 1 joule per coulomb; that is, 1 volt=1 joule per coulomb=1 newton meter per coulomb. If a unit positive charge is moved between two points, the energy required to move it is the difference in energy level between the two points and is the defined voltage. It is extremely important that the variables used to represent voltage between two points be defined in such a way that the solution will let us interpret which point is at the higher potential with respect to the other. 106 Lightning bolt Large industrial motor current Current in amperes (A) 102 Typical household appliance current 100 Causes ventricular fibrillation in humans 10–2 Human threshold of sensation 10–4 10–6 Integrated circuit (IC) memory cell current 10–8 10–10 10–12 Synaptic current (brain cell) 10–14 Figure 1.4 Typical current magnitudes. 104 3 irwin01_001-024hr.qxd CHAPTER 1 13:16 Page 4 BASIC CONCEPTS Figure 1.5 Voltage representations. A + V1=2 V – B C i r c1 u i t A + V2=–5 V – B (a) C i r c2 u i t – A V2=5 V + (b) B C i r c3 u i t (c) In Fig. 1.5a the variable that represents the voltage between points A and B has been defined as V1 , and it is assumed that point A is at a higher potential than point B, as indicated by the ± and – signs associated with the variable and defined in the figure. The ± and – signs define a reference direction for V1 . If V1 = 2 V, then the difference in potential of points A and B is 2 V and point A is at the higher potential. If a unit positive charge is moved from point A through the circuit to point B, it will give up energy to the circuit and have 2 J less energy when it reaches point B. If a unit positive charge is moved from point B to point A, extra energy must be added to the charge by the circuit, and hence the charge will end up with 2 J more energy at point A than it started with at point B. For the circuit in Fig. 1.5b, V2 = -5 V means that the potential between points A and B is 5 V and point B is at the higher potential. The voltage in Fig. 1.5b can be expressed as shown in Fig. 1.5c. In this equivalent case, the difference in potential between points A and B is V2 = 5 V, and point B is at the higher potential. Note that it is important to define a variable with a reference direction so that the answer can be interpreted to give the physical condition in the circuit. We will find that it is not possible in many cases to define the variable so that the answer is positive, and we will also find that it is not necessary to do so. As demonstrated in Figs. 1.5b and c, a negative number for a given variable, for example, V2 in Fig. 1.5b, gives exactly the same information as a positive number, that is, V2 in Fig. 1.5c, except that it has an opposite reference direction. Hence, when we define either current or voltage, it is absolutely necessary that we specify both magnitude and direction. Therefore, it is incomplete to say that the voltage between two points is 10 V or the current in a line is 2 A, since only the magnitude and not the direction for the variables has been defined. The range of magnitudes for voltage, equivalent to that for currents in Fig. 1.4, is shown in Fig. 1.6. Once again, note that this range spans many orders of magnitude. 108 Figure 1.6 Typical voltage magnitudes. 106 104 Voltage in volts (V) 4 30-06-2010 Lightning bolt High-voltage transmission lines Voltage on a TV picture tube Large industrial motors ac outlet plug in U.S. households 102 100 10–2 10–4 Car battery Voltage on integrated circuits Flashlight battery Voltage across human chest produced by the heart (EKG) Voltage between two points on human scalp (EEG) 10–6 10–8 10–10 Antenna of a radio receiver irwin01_001-024hr.qxd 30-06-2010 13:16 Page 5 SECTION 1.2 BASIC QUANTITIES 5 Figure 1.7 Flashlight circuit. Switch – Battery + Light bulb At this point we have presented the conventions that we employ in our discussions of current and voltage. Energy is yet another important term of basic significance. Let’s investigate the voltage–current relationships for energy transfer using the flashlight shown in Fig. 1.7. The basic elements of a flashlight are a battery, a switch, a light bulb, and connecting wires. Assuming a good battery, we all know that the light bulb will glow when the switch is closed. A current now flows in this closed circuit as charges flow out of the positive terminal of the battery through the switch and light bulb and back into the negative terminal of the battery. The current heats up the filament in the bulb, causing it to glow and emit light. The light bulb converts electrical energy to thermal energy; as a result, charges passing through the bulb lose energy. These charges acquire energy as they pass through the battery as chemical energy is converted to electrical energy. An energy conversion process is occurring in the flashlight as the chemical energy in the battery is converted to electrical energy, which is then converted to thermal energy in the light bulb. Figure 1.8 Flashlight circuit with voltages and current. – Battery – Vbattery + + I + Vbulb – Let’s redraw the flashlight as shown in Fig. 1.8. There is a current I flowing in this diagram. Since we know that the light bulb uses energy, the charges coming out of the bulb have less energy than those entering the light bulb. In other words, the charges expend energy as they move through the bulb. This is indicated by the voltage shown across the bulb. The charges gain energy as they pass through the battery, which is indicated by the voltage across the battery. Note the voltage–current relationships for the battery and bulb. We know that the bulb is absorbing energy; the current is entering the positive terminal of the voltage. For the battery, the current is leaving the positive terminal, which indicates that energy is being supplied. This is further illustrated in Fig. 1.9, where a circuit element has been extracted from a larger circuit for examination. In Fig. 1.9a, energy is being supplied to the element by whatever is attached to the terminals. Note that 2 A, that is, 2 C of charge are moving from point A to point B through the element each second. Each coulomb loses 3 J of energy as it passes through the element from point A to point B. Therefore, the element is absorbing 6 J of energy per second. Note that when the element is absorbing energy, a positive current enters the positive terminal. In Fig. 1.9b energy is being supplied by the element to whatever is connected to terminals A-B. In this case, note that when the element is supplying energy, a positive current enters the negative terminal and leaves via the positive terminal. In this convention, a negative current in one direction is equivalent to a positive current in the opposite direction, and vice versa. Similarly, a negative voltage in one direction is equivalent to a positive voltage in the opposite direction. A I=2 A + 3V – B I=2 A (a) A I=2 A + 3V – B I=2 A (b) Figure 1.9 Voltage–current relationships for (a) energy absorbed and (b) energy supplied. irwin01_001-024hr.qxd 6 30-06-2010 CHAPTER 1 13:16 Page 6 BASIC CONCEPTS EXAMPLE Suppose that your car will not start. To determine whether the battery is faulty, you turn on the light switch and find that the lights are very dim, indicating a weak battery. You borrow a friend’s car and a set of jumper cables. However, how do you connect his car’s battery to yours? What do you want his battery to do? 1.1 SOLUTION Essentially, his car’s battery must supply energy to yours, and therefore it should be connected in the manner shown in Fig. 1.10. Note that the positive current leaves the positive terminal of the good battery (supplying energy) and enters the positive terminal of the weak battery (absorbing energy). Note that the same connections are used when charging a battery. I Figure 1.10 Diagram for Example 1.1. I + – + – Good battery Weak battery In practical applications there are often considerations other than simply the electrical relations (e.g., safety). Such is the case with jump-starting an automobile. Automobile batteries produce explosive gases that can be ignited accidentally, causing severe physical injury. Be safe—follow the procedure described in your auto owner’s manual. We have defined voltage in joules per coulomb as the energy required to move a positive charge of 1 C through an element. If we assume that we are dealing with a differential amount of charge and energy, then v = dw dq 1.2 Multiplying this quantity by the current in the element yields vi = i(t) + v(t) dw dq dw a b = = p dq dt dt 1.3 which is the time rate of change of energy or power measured in joules per second, or watts (W). Since, in general, both v and i are functions of time, p is also a time-varying quantity. Therefore, the change in energy from time t1 to time t2 can be found by integrating Eq. (1.3); that is, – t2 ¢w = Figure 1.11 Sign convention for power. [hint] The passive sign convention is used to determine whether power is being absorbed or supplied. 3t1 p dt = t2 3t1 vi dt 1.4 At this point, let us summarize our sign convention for power. To determine the sign of any of the quantities involved, the variables for the current and voltage should be arranged as shown in Fig. 1.11. The variable for the voltage v(t) is defined as the voltage across the element with the positive reference at the same terminal that the current variable i(t) is entering. This convention is called the passive sign convention and will be so noted in the remainder of this book. The product of v and i, with their attendant signs, will determine the magnitude and sign of the power. If the sign of the power is positive, power is being absorbed by the element; if the sign is negative, power is being supplied by the element. irwin01_001-024hr.qxd 30-06-2010 13:16 Page 7 SECTION 1.2 Given the two diagrams shown in Fig. 1.12, determine whether the element is absorbing or supplying power and how much. BASIC QUANTITIES EXAMPLE 1.2 –2 A 4A – + – 2V + 2V + 2V – 2V + – Figure 1.12 (a) Elements for Example 1.2. (b) In Fig. 1.12a the power is P=(2 V)(–4 A)=–8 W. Therefore, the element is supplying power. In Fig. 1.12b, the power is P=(2 V)(–2 A)=–4 W. Therefore, the element is supplying power. SOLUTION Learning Assessment E1.1 Determine the amount of power absorbed or supplied by the elements in Fig. E1.1. + + 12 V – I=4 A V1=12 V – Figure E1.1 ANSWER: (a) P = -48 W; (b) P = 8 W. I=2 A + + 4V – V1=4 V – (b) (a) We wish to determine the unknown voltage or current in Fig. 1.13. EXAMPLE I=? 1.3 5A –A A V1=? P=–20 W P=40 W 5V ±B B – 5V + Figure 1.13 (a) (b) In Fig. 1.13a, a power of –20 W indicates that the element is delivering power. Therefore, the current enters the negative terminal (terminal A), and from Eq. (1.3) the voltage is 4 V. Thus, B is the positive terminal, A is the negative terminal, and the voltage between them is 4 V. In Fig 1.13b, a power of ±40 W indicates that the element is absorbing power and, therefore, the current should enter the positive terminal B. The current thus has a value of –8 A, as shown in the figure. Elements for Example 1.3. SOLUTION 7 irwin01_001-024hr.qxd 8 30-06-2010 CHAPTER 1 13:16 Page 8 BASIC CONCEPTS Learning Assessment E1.2 Determine the unknown variables in Fig. E1.2. ANSWER: (a) V1 = -20 V; (b) I = -5 A. I=2 A P=40 W – V1=? – + Figure E1.2 P=–50 W V1=10 V + (a) + 10 V – I=? (b) Finally, it is important to note that our electrical networks satisfy the principle of conservation of energy. Because of the relationship between energy and power, it can be implied that power is also conserved in an electrical network. This result was formally stated in 1952 by B. D. H. Tellegen and is known as Tellegen’s theorem—the sum of the powers absorbed by all elements in an electrical network is zero. Another statement of this theorem is that the power supplied in a network is exactly equal to the power absorbed. Checking to verify that Tellegen’s theorem is satisfied for a particular network is one way to check our calculations when analyzing electrical networks. 1.3 Circuit Elements Thus far we have defined voltage, current, and power. In the remainder of this chapter we will define both independent and dependent current and voltage sources. Although we will assume ideal elements, we will try to indicate the shortcomings of these assumptions as we proceed with the discussion. In general, the elements we will define are terminal devices that are completely characterized by the current through the element and/or the voltage across it. These elements, which we will employ in constructing electric circuits, will be broadly classified as being either active or passive. The distinction between these two classifications depends essentially on one thing—whether they supply or absorb energy. As the words themselves imply, an active element is capable of generating energy and a passive element cannot generate energy. However, later we will show that some passive elements are capable of storing energy. Typical active elements are batteries and generators. The three common passive elements are resistors, capacitors, and inductors. In Chapter 2 we will launch an examination of passive elements by discussing the resistor in detail. Before proceeding with that element, we first present some very important active elements. 1. Independent voltage source 2. Independent current source 3. Two dependent voltage sources 4. Two dependent current sources I N D E P E N D E N T S O U R C E S An independent voltage source is a two-terminal element that maintains a specified voltage between its terminals regardless of the current through it as shown by the v-i plot in Fig. 1.14a. The general symbol for an independent source, a circle, is also shown in Fig. 1.14a. As the figure indicates, terminal A is v(t) volts positive with respect to terminal B. In contrast to the independent voltage source, the independent current source is a twoterminal element that maintains a specified current regardless of the voltage across its terminals, as illustrated by the v-i plot in Fig. 1.14b. The general symbol for an independent current source is also shown in Fig. 1.14b, where i(t) is the specified current and the arrow indicates the positive direction of current flow. irwin01_001-024hr.qxd 30-06-2010 13:16 Page 9 SECTION 1.3 CIRCUIT ELEMENTS v v 9 Figure 1.14 Symbols for (a) independent voltage source, (b) independent current source. i i A A v(t) + – i(t) B B (a) (b) In their normal mode of operation, independent sources supply power to the remainder of the circuit. However, they may also be connected into a circuit in such a way that they absorb power. A simple example of this latter case is a battery-charging circuit such as that shown in Example 1.1. It is important that we pause here to interject a comment concerning a shortcoming of the models. In general, mathematical models approximate actual physical systems only under a certain range of conditions. Rarely does a model accurately represent a physical system under every set of conditions. To illustrate this point, consider the model for the voltage source in Fig. 1.14a. We assume that the voltage source delivers v volts regardless of what is connected to its terminals. Theoretically, we could adjust the external circuit so that an infinite amount of current would flow, and therefore the voltage source would deliver an infinite amount of power. This is, of course, physically impossible. A similar argument could be made for the independent current source. Hence, the reader is cautioned to keep in mind that models have limitations and thus are valid representations of physical systems only under certain conditions. For example, can the independent voltage source be utilized to model the battery in an automobile under all operating conditions? With the headlights on, turn on the radio. Do the headlights dim with the radio on? They probably won’t if the sound system in your automobile was installed at the factory. If you try to crank your car with the headlights on, you will notice that the lights dim. The starter in your car draws considerable current, thus causing the voltage at the battery terminals to drop and dimming the headlights. The independent voltage source is a good model for the battery with the radio turned on; however, an improved model is needed for your battery to predict its performance under cranking conditions. Determine the power absorbed or supplied by the elements in the network in Fig. 1.15. I=2 A + 24 V 6V EXAMPLE 1.4 – 1 I=2 A + 2 18 V – + – I=2 A Figure 1.15 Network for Example 1.4. irwin01_001-024hr.qxd 10 30-06-2010 CHAPTER 1 13:16 Page 10 BASIC CONCEPTS SOLUTION [hint] Elements that are connected in series have the same current. The current flow is out of the positive terminal of the 24-V source, and therefore this element is supplying (2)(24)=48 W of power. The current is into the positive terminals of elements 1 and 2, and therefore elements 1 and 2 are absorbing (2)(6)=12 W and (2)(18)=36 W, respectively. Note that the power supplied is equal to the power absorbed. Learning Assessment E1.3 Find the power that is absorbed or supplied by the elements in Fig. E1.3. 18 V I=3 A + – 1 I=3 A + 3A 12 V ANSWER: Current source supplies 36 W, element 1 absorbs 54 W, and element 2 supplies 18 W. 2 – – 6V + Figure E1.3 D E P E N D E N T S O U R C E S In contrast to the independent sources, which produce a particular voltage or current completely unaffected by what is happening in the remainder of the circuit, dependent sources generate a voltage or current that is determined by a voltage or current at a specified location in the circuit. These sources are very important because they are an integral part of the mathematical models used to describe the behavior of many electronic circuit elements. For example, metal-oxide-semiconductor field-effect transistors (MOSFETs) and bipolar transistors, both of which are commonly found in a host of electronic equipment, are modeled with dependent sources, and therefore the analysis of electronic circuits involves the use of these controlled elements. In contrast to the circle used to represent independent sources, a diamond is used to represent a dependent or controlled source. Fig. 1.16 illustrates the four types of dependent sources. The input terminals on the left represent the voltage or current that controls the dependent source, and the output terminals on the right represent the output current or voltage of the controlled source. Note that in Figs. 1.16a and d, the quantities and  are dimensionless constants because we are transforming voltage to voltage and current to current. This is not the case in Figs. 1.16b and c; hence, when we employ these elements a short time later, we must describe the units of the factors r and g. Figure 1.16 Four different types of dependent sources. iS + vS + v=vS – + v=riS – – (a) (b) iS + i=gvS vS i=iS – (c) (d) irwin01_001-024hr.qxd 30-06-2010 13:16 Page 11 SECTION 1.3 Given the two networks shown in Fig. 1.17, we wish to determine the outputs. + 11 EXAMPLE In Fig. 1.17a the output voltage is Vo = VS or Vo = 20 VS=(20)(2 V)=40 V. Note that the output voltage has been amplified from 2 V at the input terminals to 40 V at the output terminals; that is, the circuit is a voltage amplifier with an amplification factor of 20. IS=1 mA CIRCUIT ELEMENTS SOLUTION 1.5 Io + + – VS=2 V Vo 20VS=Vo – 50IS=Io – Figure 1.17 (a) Circuits for Example 1.5. (b) In Fig. 1.17b, the output current is Io = IS = (50)(1 mA) = 50 mA; that is, the circuit has a current gain of 50, meaning that the output current is 50 times greater than the input current. Learning Assessment E1.4 Determine the power supplied by the dependent sources in Fig. E1.4. ANSWER: (a) Power supplied = 80 W; (b) power supplied = 160 W. IS=4 A Io=2 A + + ± – VS=4 V 10VS 4 IS 10 V 1 1 – – (a) (b) Figure E1.4 Calculate the power absorbed by each element in the network of Fig. 1.18. Also verify that Tellegen’s theorem is satisfied by this network. + 1A + 8V 4 EXAMPLE 1.6 12 V 3 + - 4V 2 1A 2A + 24 V 3A ± – 16 V 12 V 1 - ± – Figure 1.18 1A 2A Let’s calculate the power absorbed by each element using the sign convention for power. P1 = (16)(1) = 16 W P2 = (4)(1) = 4 W P3 = (12)(1) = 12 W Circuit used in Example 1.6. SOLUTION irwin01_001-024hr.qxd 12 30-06-2010 CHAPTER 1 13:16 Page 12 BASIC CONCEPTS P4 = (8)(2) = 16 W P12V = (12)(2) = 24 W P24V = (24)(-3) = -72 W Note that to calculate the power absorbed by the 24-V source, the current of 3 A flowing up through the source was changed to a current ⫺3 A flowing down through the 24-V source. Let’s sum up the power absorbed by all elements: 16 ⫹ 4 ⫹ 12 ⫹ 16 ⫹ 24 ⫺ 72 ⫽ 0 This sum is zero, which verifies that Tellegen’s theorem is satisfied. EXAMPLE Use Tellegen’s theorem to find the current Io in the network in Fig. 1.19. 1.7 – 6V + Ix=2 A 2A + 6V Io – Figure 1.19 3A 12 V 9A + 3 10 V – Circuit used in Example 1.7. SOLUTION – 1 ± – + 2 4V 11 A ± – 8Ix 8A First, we must determine the power absorbed by each element in the network. Using the sign convention for power, we find P2 A = (6)(-2) = -12 W P1 = (6)AIo B = 6Io W P2 = (12)(-9) = -108 W P3 = (10)(-3) = -30 W P4 V = (4)(-8) = -32 W PDS = A8Ix B(11) = (16)(11) = 176 W Applying Tellegen’s theorem yields -12 + 6Io - 108 - 30 - 32 + 176 = 0 or 6Io + 176 = 12 + 108 + 30 + 32 Hence, Io = 1A Learning Assessment E1.5 Find the power that is absorbed or supplied by the circuit elements in the network in Fig. E1.5. + 8V – 1 24 V Figure E1.5 ± – ± – Ix=4 A 4Ix ANSWER: P24 V = 96 W supplied; P1 = 32 W absorbed; P4Ix = 64 W absorbed. irwin01_001-024hr.qxd 30-06-2010 13:16 Page 13 SECTION 1.3 E1.6 Find the power that is absorbed 2 Ix or supplied by the network elements in Fig. E1.6. –+ 24 V ANSWER: P24V = 36 W supplied, P12V = 18 W absorbed, P21x = 4.5 W supplied, P1 = 9 W absorbed, P2 = 13.5 W absorbed. 1 + – 6V + – + – – 9V + 12 V Ix=1.5 A 2 Figure E1.6 E1.7 Find Ix in Fig. E1.7 using CIRCUIT ELEMENTS 1A 2A ANSWER: Ix = -2 A. Tellegen’s theorem. + + Ix 10 V + 1 25 V 3 – 15 V – 5A – + 2 15 V – + – 10 V Figure E1.7 The charge that enters the BOX is shown in Fig. 1.20. Calculate and sketch the current flowing into and the power absorbed by the BOX between 0 and 10 milliseconds. EXAMPLE 1.8 i (t) 12 V ± – BOX q(t) (mC) 3 2 1 5 1 2 3 4 6 7 8 9 10 t (ms) –1 –2 –3 Figure 1.20 Diagrams for Example 1.8. 13 irwin01_001-024hr.qxd 14 30-06-2010 CHAPTER 1 13:16 Page 14 BASIC CONCEPTS SOLUTION Recall that current is related to charge by i(t) = dq(t) dt . The current is equal to the slope of the charge waveform. 0 ⱕ t ⱕ 1 ms i(t) = 0 -3 i(t) = -3 3 * 10 - 1 * 10 = 2A 2 * 10-3 - 1 * 10-3 1 ⱕ t ⱕ 2 ms 2 ⱕ t ⱕ 3 ms i(t) = 0 i(t) = -2 * 10-3 - 3 * 10-3 = -2.5 A 5 * 10-3 - 3 * 10-3 3 ⱕ t ⱕ 5 ms 5 ⱕ t ⱕ 6 ms i(t) = 0 -3 i(t) = -3 - (-2 * 10 ) 2 * 10 -3 9 * 10 6 ⱕ t ⱕ 9 ms = 1.33 A - 6 * 10-3 t ⱖ 9 ms i(t) = 0 The current is plotted with the charge waveform in Fig. 1.21. Note that the current is zero during times when the charge is a constant value. When the charge is increasing, the current is positive, and when the charge is decreasing, the current is negative. Figure 1.21 Charge and current waveforms for Example 1.8. q(t) (mC), i(t) (A) 3 2 1 5 1 2 3 4 6 7 8 9 10 t (ms) –1 –2 –3 The power absorbed by the BOX is 12 ⭈ i(t). p(t) p(t) p(t) p(t) p(t) p(t) = = = = = = 12*0 = 0 12*2 = 24 W 12*0 = 0 12*(-2.5) = - 30 W 12*0 = 0 12*1.33 = 16 W p(t) = 12*0 = 0 0 ⱕ t ⱕ 1 ms 1 ⱕ t ⱕ 2 ms 2 ⱕ t ⱕ 3 ms 3 ⱕ t ⱕ 5 ms 5 ⱕ t ⱕ 6 ms 6 ⱕ t ⱕ 9 ms t ⱖ 9 ms The power absorbed by the BOX is plotted in Fig. 1.22. For the time intervals, 1 ⱕ t ⱕ 2 ms and 6 ⱕ t ⱕ 9 ms, the BOX is absorbing power. During the time interval 3 ⱕ t ⱕ 5 ms, the power absorbed by the BOX is negative, which indicates that the BOX is supplying power to the 12-V source. irwin01_001-024hr.qxd 30-06-2010 13:16 Page 15 SECTION 1.3 CIRCUIT ELEMENTS Figure 1.22 p(t) (W) Power waveform for Example 1.8. 36 24 12 5 1 2 3 6 4 7 8 9 10 t (ms) –12 –24 –36 Learning Assessment E1.8 The power absorbed by the BOX in Fig. El.8 is p(t) = 2.5e-4t W. Compute the energy and charge delivered to the BOX in the time interval 0 6 t 6 250 ms. i (t) 50e–t V ANSWER: 395.1 mJ, 8.8 mC. + – BOX Figure E1.8 E1.9 The energy absorbed by the BOX in Fig. El.9 is given below. Calculate and sketch the current flowing into the BOX. Also calculate the charge that enters the BOX between 0 and 12 seconds. i (t) 10 V + – BOX w(t) (J) 5 7 1 –2.5 Figure E1.9 2 3 4 5 6 8 9 10 11 12 t (s) 15 irwin01_001-024hr.qxd 16 30-06-2010 CHAPTER 1 13:16 Page 16 BASIC CONCEPTS ANSWER: Q = 0. i(t) (A) 0.25 0.125 4 1 2 3 5 6 7 8 9 10 11 12 t (s) –0.125 –0.25 EXAMPLE 1.9 Figure 1.23 Charging a Motorola RAZR® and Apple iPod® from USB ports. (Courtesy of Mark Nelms and Jo Ann Loden) A Universal Serial Bus (USB) port is a common feature on both desktop and notebook computers as well as many handheld devices such as MP3 players, digital cameras, and cell phones. The USB 2.0 specification (www.usb.org) permits data transfer between a computer and a peripheral device at rates up to 480 megabits per second. One important feature of USB is the ability to swap peripherals without having to power down a computer. USB ports are also capable of supplying power to external peripherals. Fig. 1.23 shows a Motorola RAZR® and an Apple iPod® being charged from the USB ports on a notebook computer. A USB cable is a four-conductor cable with two signal conductors and two conductors for providing power. The amount of current that can be provided over a USB port is defined in the USB specification in terms of unit loads, where one unit load is specified to be 100 mA. All USB ports default to low-power ports at one unit load, but can be changed under software control to high-power ports capable of supplying up to five unit loads or 500 mA. irwin01_001-024hr.qxd 30-06-2010 13:16 Page 17 PROBLEMS 1. 2. 1. 2. • A 680 mAh lithium-ion battery is standard in a Motorola RAZR®. If this battery is completely discharged (i.e., 0 mAh), how long will it take to recharge the battery to its full capacity of 680 mAh from a low-power USB port? How much charge is stored in the battery at the end of the charging process? A third-generation iPod® with a 630 mAh lithium-ion battery is to be recharged from a high-power USB port supplying 150 mA of current. At the beginning of the recharge, 7.8 C of charge are stored in the battery. The recharging process halts when the stored charge reaches 35.9 C. How long does it take to recharge the battery? A low-power USB port operates at 100 mA. Assuming that the charging current from the USB port remains at 100 mA throughout the charging process, the time required to recharge the battery is 680 mAh兾100 mA = 6.8 h. The charge stored in the battery when fully charged is 680mAh ⴢ 60 s兾h = 40,800 mAs = 40.8 As = 40.8 C. The charge supplied to the battery during the recharging process is 35.9 - 7.8 = 28.1 C. This corresponds to 28.1 As = 28,100 mAs ⴢ 1h兾60s = 468.3 mAh. Assuming a constant charging current of 150 mA from the high-power USB port, the time required to recharge the battery is 468.3 mAh兾150 mA = 3.12 h. SOLUTION SUMMARY ■ ■ The standard prefixes employed p = 10-12 k = 103 n = 10-9 = 10-6 M = 106 m = 10-3 T = 1012 ■ dq(t) dt The passive sign convention The passive sign convention states that if the voltage and current associated with an element are as shown in Fig. 1.11, the product of v and i, with their attendant signs, determines the magnitude and sign of the power. If the sign is positive, power is being absorbed by the element, and if the sign is negative, the element is supplying power. ■ Independent and dependent sources An ideal independent voltage (current) source is a two-terminal element that maintains a specified voltage (current) between its terminals, regardless of the current (voltage) through (across) the element. Dependent or controlled sources generate a voltage or current that is determined by a voltage or current at a specified location in the circuit. ■ Conservation of energy The electric circuits under investigation satisfy the conservation of energy. ■ Tellegen’s theorem The sum of the powers absorbed by all elements in an electrical network is zero. t or q(t) = 3-q i(x) dx The relationships among power, energy, current, and voltage dw p = = vi dt t2 ¢w = ■ G = 109 The relationships between current and charge i(t) = • 17 3t1 t2 p dt = 3t1 vi dt PROBLEMS 1.1 If the current in an electric conductor is 2.4 A, how many coulombs of charge pass any point in a 30-second interval? 1.2 Determine the time interval required for a 12-A battery charger to deliver 4800 C. 1.3 A lightning bolt carrying 30,000 A lasts for 50 microseconds. If the lightning strikes an airplane flying at 20,000 feet, what is the charge deposited on the plane? 1.4 If a 12-V battery delivers 100 J in 5 s, find (a) the amount of charge delivered and (b) the current produced. irwin01_001-024hr.qxd 18 30-06-2010 CHAPTER 1 13:16 Page 18 BASIC CONCEPTS 1.5 The current in a conductor is 1.5 A. How many coulombs of 1.11 The charge entering the positive terminal of an element is given by the expression q(t) = -12e-2t mC. The power delivered to the element is p(t) = 2.4e-3t W. Compute the current in the element, the voltage across the element, and the energy delivered to the element in the time interval 0 6 t 6 100 ms. charge pass any point in a time interval of 1.5 minutes? 1.6 If 60 C of charge pass through an electric conductor in 30 seconds, determine the current in the conductor. 1.7 Determine the number of coulombs of charge produced by a 12-A battery charger in an hour. 1.12 The voltage across an element is 12e-2t V . The current 1.8 Five coulombs of charge pass through the element in Fig. P1.8 from point A to point B. If the energy absorbed by the element is 120 J, determine the voltage across the element. B entering the positive terminal of the element is 2e-2t A. Find the energy absorbed by the element in 1.5 s starting from t ⫽ 0. 1.13 The power absorbed by the BOX in Fig. P1.13 is 2e-2t W. Calculate the amount of charge that enters the BOX between 0.1 and 0.4 seconds. + V1 - A 4e–t V Figure P1.8 + – BOX 1.9 The current that enters an element is shown in Fig. P1.9. Find the charge that enters the element in the time interval 0 6 t 6 20 s. Figure P1.13 1.14 The power absorbed by the BOX in Fig. P1.14 is i(t) mA 0.1e-4t W. Calculate the energy absorbed by the BOX during this same time interval. 10 0 10 20 t (s) 10e–2t V Figure P1.9 + – BOX 1.10 The charge entering the positive terminal of an element is q(t) = -30e-4t mC. If the voltage across the element is 120e-2t V, determine the energy delivered to the element in the time interval 0 6 t 6 50 ms. Figure P1.14 1.15 The energy absorbed by the BOX in Fig. P1.15 is shown below. How much charge enters the BOX between 0 and 10 milliseconds? w(t) (mJ) 15 i (t) 10 15 V + – BOX 5 1 –5 –10 –15 Figure P1.15 2 3 4 5 6 7 8 9 10 t (ms) irwin01_001-024hr.qxd 30-06-2010 13:16 Page 19 PROBLEMS 19 1.16 The charge that enters the BOX in Fig. P1.16 is shown in the graph below. Calculate and sketch the current flowing into and the power absorbed by the BOX between 0 and 10 milliseconds. i (t) + – 12 V BOX q(t) (mC) 3 2 1 5 1 2 3 4 6 7 8 9 10 t (ms) –1 –2 –3 Figure P1.16 1.17 The energy absorbed by the BOX in Fig. P1.17 is given below. Calculate and sketch the current flowing into the BOX. Also calculate the charge which enters the BOX between 0 and 12 seconds. i (t) 10 V + – BOX w(t) (J) 5 6 1 –2.5 Figure P1.17 2 3 4 5 7 8 10 9 12 11 t (s) irwin01_001-024hr.qxd 20 30-06-2010 CHAPTER 1 13:16 Page 20 BASIC CONCEPTS 1.18 The charge entering the upper terminal of the BOX in Fig. P1.18 is shown below. How much energy is absorbed by the BOX between 0 and 9 seconds? i (t) 12 V + – BOX q(t) (C) 1 0.5 1 2 3 4 5 6 7 8 9 t (s) –0.5 –1 –1.5 Figure P1.18 1.19 The energy absorbed by the BOX in Fig. P1.19 is shown in the graph below. Calculate and sketch the current flowing into the BOX between 0 and 10 milliseconds. i (t) 12 V + – BOX w(t) (mJ) 30 20 10 5 1 –10 –20 –30 Figure P1.19 2 3 4 6 7 8 9 10 t (ms) irwin01_001-024hr.qxd 30-06-2010 13:16 Page 21 PROBLEMS 1.20 Determine the amount of power absorbed or supplied by the element in Fig. P1.20 if 21 1.24 Element B in the diagram in Fig. P1.24 supplies 60 W of power. Calculate Ix. (a) V1 = 9 V and I = 2A (b) V1 = 9 V and I = -3A (c) V1 = -12 V and I = 2A – 24 V B (d) V1 = -12 V and I = -3A + + I Ix V1 Figure P1.24 - 1.25 Element B in the diagram in Fig. P1.25 supplies 72 W of power. Calculate VA. Figure P1.20 3A 1.21 Calculate the power absorbed by element A in Fig. P1.21. + VA 3A B – – 15 V A + Figure P1.25 1.26 Element B in the diagram in Fig. P1.26 supplies 72 W of power. Calculate Ix. Figure P1.21 1.22 Calculate the power supplied by element A in Fig. P1.22. + 2A 18 V B + – 20 V A – Ix Figure P1.26 1.27 (a) In Fig. P1.27 (a), P1 = 36 W. Is element 2 absorbing Figure P1.22 or supplying power, and how much? 1.23 Element A in the diagram in Fig. P1.23 absorbs 30 W of power. Calculate Vx. (b) In Fig. P1.27 (b), P2 = -48 W. Is element 1 absorb- ing or supplying power, and how much? 2A + 12 V – + 2 6V - Vx A – (a) Figure P1.23 6V + + 2 24 V - 1 + Figure P1.27 1 (b) irwin01_001-024hr.qxd 22 30-06-2010 CHAPTER 1 13:16 Page 22 BASIC CONCEPTS 1.28 Two elements are connected in series, as shown in Fig. P1.28. Element 1 supplies 24 W of power. Is element 2 absorbing or supplying power, and how much? 1.32 Find the power that is absorbed or supplied by the network elements in Fig. P1.32. Ix=2 A + 1 3V – 2 6V + + 8V - 2A 1 ± – ± – 12 V 2 Ix 2A Figure P1.28 (a) 1.29 Element 2 in Fig. P1.29 absorbed 32 W. Find the power 24 V absorbed or supplied by elements 1 and 3. 2A 4V + + 2 8V – + 3 12 V 1 ± – –± 20 V + 1 2A 4Ix Ix=2 A + 2 12 V - 2A (b) Figure P1.32 Figure P1.29 1.33 Compute the power that is absorbed or supplied by the elements in the network in Fig. P1.33. 1.30 Choose Is such that the power absorbed by element 2 in Fig. P1.30 is 7 W. + 4V Ix=4 A – 1 + + Is 6V – 36 V 2 2V – 1.31 Find the power that is absorbed or supplied by the circuit elements in Fig. P1.31. 6V ± – 2 + 24 V - + 3 28 V - in Fig. P1.34. - 2A 2A+ 4V 1 2Vx – –+ + + - 2A - 14 V 12 V + – 2 – Figure P1.34 (a) Ix=4 A + 8V 1.35 Find Ix in the network in Fig. P1.35. - 1 1Ix Ix +12 V – + - 4A –+ 1 2Ix 4A 2A 36 V + – (b) Figure P1.31 Vx 2A 2A + 16 V - 2A 1.34 Find the power that is absorbed or supplied by element 2 1 + 20 V –± Figure P1.33 Figure P1.30 + 1Ix 12 V + 1 2A Figure P1.35 2 + 24 V – 2A 3 + 28 V – irwin01_001-024hr.qxd 30-06-2010 13:16 Page 23 23 PROBLEMS 1.36 Determine the power absorbed by element 1 in Fig. P1.36. 1.40 Find Vx in the network in Fig. P1.40 using Tellegen’s theorem. Ix +12 V – 1 + – 2 + 3 16 V – 2Ix 24 V – 2A 2A + + – 36 V 8V 9V + + 1 + 12 V 16 V + 3 2 + 24 V Vx - + 12 V Figure P1.36 Figure P1.40 1.37 Find the power absorbed or supplied by element 1 in Fig. P1.37. theorem. 6V 1 – Ix + 4V 2 – 2A + – Ix 24 V + – 2Ix 2 A+ ± – – Fig. P1.38. –+ Ix 4A + 3 2A 2A – Vx + 4 20 V – 9A theorem. 8V 1 4A + 4Ix 12 V – Figure P1.39 ± – 12 V + 2 8V – 4A 4A 8V + Figure P1.42 24 V + 6A - 1.43 Find Io in the network in Fig. P1.43 using Tellegen’s 2A P1.39. –+ 16 V 3A 6 A+ – + + 1.39 Find the power absorbed or supplied by element 1 in Fig. 1 VS - + Figure P1.38 4V + 6V - 2A 9A 10 V 16 V + 2Vx – 6V 3A + – + Ix 1.42 Is the source Vs in the network in Fig. P1.42 absorbing + 2 12 V 24 V 12 V 18 V + - - + 12 V - 12 V 2A + – 8V or supplying power, and how much? – 1 + Figure P1.41 1.38 Find the power absorbed or supplied by element 3 in 4V - 2A + 20 V Figure P1.37 + 4V ±– + 18 V 1.41 Find Ix in the circuit in Fig. P1.41 using Tellegen’s + + 3 4 20 V – 2A Ix 20 V – 6V 3 2 10 V + Io 4Ix 3A ± – - Figure P1.43 8V 6 + 3A + 5 6V 1A Ix=2 A + 4 16 V - irwin01_001-024hr.qxd 24 30-06-2010 CHAPTER 1 13:16 Page 24 BASIC CONCEPTS 1.44 Calculate the power absorbed by each element in the circuit in Fig. P1.44. Also verify Tellegen’s theorem is satisfied by this circuit. 3Ix 1.46 In the circuit in Fig. P1.46, element 1 absorbs 40 W, element 2 supplies 50 W, element 3 supplies 25 W, and element 4 absorbs 15 W. How much power is supplied by element 5? 24 V + – 5 2A –+ 1 3 2A 12 V 6V 9V + – + – + 1 2 4 2A 4A 4A + + 24 V + 15 V + 12 V 6A 6V 3 – – – – Ix = 2 A 4A Figure P1.44 1.45 Calculate the power absorbed by each element in the circuit in Fig. P1.45. Also verify that Tellegen’s theorem is satisfied by this circuit. 10 V + – 3 4A + 5V – + 2 5V + – - 5V 5A 1 + Figure P1.45 – 3A 1A 40 V 4 15 V + 4A 30 V – + – + 1A 5 10 V – Figure P1.46 2 4 5 irwin02_025-100hr.qxd 30-06-2010 13:14 Page 25 C H A PT E R RESISTIVE CIRCUITS 2 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Be able to use Ohm’s law to solve electric circuits ■ Be able to apply Kirchhoff’s current law and Kirchhoff’s voltage law to solve electric circuits ■ Know how to analyze single-loop and single- node-pair circuits ■ Know how to combine resistors in series and parallel ■ Be able to use voltage and current division to solve simple electric circuits ■ Understand when and how to apply wye-delta transformations in the analysis of electric circuits ■ Know how to analyze electric circuits contain- Courtesy of Tesla Motors T Tesla Roadster Green technologies come in many colors. The 2010 Tesla Roadster, for example, comes in Fusion Red, Arctic White, Racing Green and Electric Blue, to name a few. ing dependent sources bility. Handling qualities may be highly important to some, cost and efficiency to others. As a student of circuit analysis, you will make trade-offs in An environmentally friendly sports car that seats two, this choosing between methods of analysis for different circuit convertible has rocket acceleration and hugs the road like a topologies. This chapter describes fundamental laws that dream; it’s the world’s first high-performance electric car. apply to all circuits regardless of their complexity. Ohm’s law The Roadster contains over 6,800 safe, rechargeable lithium- governs the most common relationship between voltage and ion batteries that weigh about 1,000 pounds in total. It is current for circuits that are linear. Circuits having a single twice as efficient as hybrid cars that combine a gasoline power source with resistances having the same currents and engine and an electric motor to provide propulsion, but its others having the same voltage will be analyzed using the fantastic performance comes at a cost of over $100,000. series-parallel method. You’ll learn more techniques in the Choosing between an all-electric vehicle and a hybrid requires trade-offs on a wide range of criteria: performance, chapters that follow, as you begin to master the same principles used by the designers of the Tesla Roadster. cost, efficiency, effects on the environment, safety, and relia- 25 irwin02_025-100hr.qxd 26 22-07-2010 CHAPTER 2 15:03 Page 26 RESISTIVE CIRCUITS 2.1 Ohm’s law is named for the German physicist Georg Simon Ohm, who is credited with establishing the voltage–current relationship for resistance. As a result of his pioneering work, the unit of resistance bears his name. Ohm’s law states that the voltage across a resistance is directly proportional to the current flowing through it. The resistance, measured in ohms, is the constant of proportionality between the voltage and current. A circuit element whose electrical characteristic is primarily resistive is called a resistor and is represented by the symbol shown in Fig. 2.1a. A resistor is a physical device that can be purchased in certain standard values in an electronic parts store. These resistors, which find use in a variety of electrical applications, are normally carbon composition or wirewound. In addition, resistors can be fabricated using thick oxide or thin metal films for use in hybrid circuits, or they can be diffused in semiconductor integrated circuits. Some typical discrete resistors are shown in Fig. 2.1b. The mathematical relationship of Ohm’s law is illustrated by the equation Ohm’s Law [hint] The passive sign convention will be employed in conjunction with Ohm’s law. v(t)=R i(t), where R ⭌ 0 2.1 or equivalently, by the voltage–current characteristic shown in Fig. 2.2a. Note carefully the relationship between the polarity of the voltage and the direction of the current. In addition, note that we have tacitly assumed that the resistor has a constant value and therefore that the voltage–current characteristic is linear. The symbol ⍀ is used to represent ohms, and therefore, 1 ⍀=1 V/A Although in our analysis we will always assume that the resistors are linear and are thus described by a straight-line characteristic that passes through the origin, it is important that readers realize that some very useful and practical elements do exist that exhibit a nonlinear resistance characteristic; that is, the voltage–current relationship is not a straight line. Figure 2.1 (a) Symbol for a resistor; (b) some practical devices. (1), (2), and (3) are highpower resistors. (4) and (5) are high-wattage fixed resistors. (6) is a highprecision resistor. (7)–(12) are fixed resistors with different power ratings. (Photo courtesy of Mark Nelms and Jo Ann Loden) i(t) + v(t) R – (a) (b) irwin02_025-100hr.qxd 30-06-2010 13:14 Page 27 SECTION 2.1 v(t) 27 Figure 2.2 v(t) Graphical representation of the voltage–current relationship for (a) a linear resistor and (b) a light bulb. R 1 i(t) i(t) (a) (b) The light bulb from the flashlight in Chapter 1 is an example of an element that exhibits a nonlinear characteristic. A typical characteristic for a light bulb is shown in Fig. 2.2b. Since a resistor is a passive element, the proper current–voltage relationship is illustrated in Fig. 2.1a. The power supplied to the terminals is absorbed by the resistor. Note that the charge moves from the higher to the lower potential as it passes through the resistor and the energy absorbed is dissipated by the resistor in the form of heat. As indicated in Chapter 1, the rate of energy dissipation is the instantaneous power, and therefore p(t)=v(t)i(t) 2.2 which, using Eq. (2.1), can be written as p(t) = Ri2(t) = v2(t) R 2.3 This equation illustrates that the power is a nonlinear function of either current or voltage and that it is always a positive quantity. Conductance, represented by the symbol G, is another quantity with wide application in circuit analysis. By definition, conductance is the reciprocal of resistance; that is, G = 1 R 2.4 The unit of conductance is the siemens, and the relationship between units is 1 S=1 A/ V Using Eq. (2.4), we can write two additional expressions, i(t)=Gv(t) 2.5 and p(t) = OHM’S LAW i2(t) G = Gv2(t) 2.6 Eq. (2.5) is another expression of Ohm’s law. Two specific values of resistance, and therefore conductance, are very important: R=0 and R=q. In examining the two cases, consider the network in Fig. 2.3a. The variable resistance symbol is used to describe a resistor such as the volume control on a radio or television set. irwin02_025-100hr.qxd 28 30-06-2010 CHAPTER 2 13:14 Page 28 RESISTIVE CIRCUITS Figure 2.3 Short-circuit and open-circuit descriptions. i(t) i(t) + v(t) R i(t) + + v(t) v(t) – – – (a) (b) (c) As the resistance is decreased and becomes smaller and smaller, we finally reach a point where the resistance is zero and the circuit is reduced to that shown in Fig. 2.3b; that is, the resistance can be replaced by a short circuit. On the other hand, if the resistance is increased and becomes larger and larger, we finally reach a point where it is essentially infinite and the resistance can be replaced by an open circuit, as shown in Fig. 2.3c. Note that in the case of a short circuit where R 0, v(t) = Ri(t) = 0 Therefore, v(t) = 0, although the current could theoretically be any value. In the opencircuit case where R = q, i(t) = v(t)兾R = 0 Therefore, the current is zero regardless of the value of the voltage across the open terminals. EXAMPLE 2.1 SOLUTION In the circuit in Fig. 2.4a, determine the current and the power absorbed by the resistor. Using Eq. (2.1), we find the current to be I=V/R=12/2k=6 mA Note that because many of the resistors employed in our analysis are in k, we will use k in the equations in place of 1000. The power absorbed by the resistor is given by Eq. (2.2) or (2.3) as P = VI = (12)A6 * 10-3 B = 0.072 W = I 2R = A6 * 10-3 B (2k) = 0.072 W 2 = V2兾R = (12)2兾2k = 0.072 W Figure 2.4 Circuits for Examples 2.1 to 2.4. I 12 V ± – I (a) (b) I=0.5 mA VS ± – 10 k P=3.6 mW ± VS – 2 k + G=50 S VS I 4 mA (c) (d) P=80 mW R irwin02_025-100hr.qxd 30-06-2010 13:14 Page 29 SECTION 2.1 The power absorbed by the 10-k resistor in Fig. 2.4b is 3.6 mW. Determine the voltage and the current in the circuit. Using the power relationship, we can determine either of the unknowns: OHM’S LAW 29 EXAMPLE 2.2 SOLUTION V 2S兾R = P V 2S = (3.6 * 10-3)(10k) VS = 6 V and I 2R = P I 2 = (3.6 * 10-3)兾10k I = 0.6 mA Furthermore, once VS is determined, I could be obtained by Ohm’s law, and likewise once I is known, then Ohm’s law could be used to derive the value of VS. Note carefully that the equations for power involve the terms I2 and V2S. Therefore, I = -0.6 mA and VS = -6 V also satisfy the mathematical equations and, in this case, the direction of both the voltage and current is reversed. Given the circuit in Fig. 2.4c, we wish to find the value of the voltage source and the power absorbed by the resistance. EXAMPLE 2.3 SOLUTION The voltage is VS = I兾G = (0.5 * 10-3)兾(50 * 10-6) = 10 V The power absorbed is then P = I 2兾G = (0.5 * 10-3)2兾(50 * 10-6) = 5 mW Or we could simply note that R = 1兾G = 20 k and therefore VS = IR = (0.5 * 10-3)(20k) = 10 V and the power could be determined using P = I 2R = V 2S兾R = VS I. EXAMPLE Given the network in Fig. 2.4d, we wish to find R and VS . SOLUTION Using the power relationship, we find that R = P兾I 2 = A80 * 10-3 B兾A4 * 10-3 B = 5 k 2 The voltage can now be derived using Ohm’s law as VS = IR = A4 * 10-3 B(5k) = 20 V The voltage could also be obtained from the remaining power relationships in Eqs. (2.2) and (2.3). 2.4 irwin02_025-100hr.qxd 30 30-06-2010 CHAPTER 2 13:14 Page 30 RESISTIVE CIRCUITS Before leaving this initial discussion of circuits containing sources and a single resistor, it is important to note a phenomenon that we will find to be true in circuits containing many sources and resistors. The presence of a voltage source between a pair of terminals tells us precisely what the voltage is between the two terminals regardless of what is happening in the balance of the network. What we do not know is the current in the voltage source. We must apply circuit analysis to the entire network to determine this current. Likewise, the presence of a current source connected between two terminals specifies the exact value of the current through the source between the terminals. What we do not know is the value of the voltage across the current source. This value must be calculated by applying circuit analysis to the entire network. Furthermore, it is worth emphasizing that when applying Ohm’s law, the relationship V = IR specifies a relationship between the voltage directly across a resistor R and the current that is present in this resistor. Ohm’s law does not apply when the voltage is present in one part of the network and the current exists in another. This is a common mistake made by students who try to apply V = IR to a resistor R in the middle of the network while using a V at some other location in the network. Learning Assessments E2.1 Given the circuits in Fig. E2.1, find (a) the current I and the power absorbed by the resistor in Fig. E2.1a, and (b) the voltage across the current source and the power supplied by the source in Fig. E2.1b. I 12 V + ± – Figure E2.1 40 k VS 0.6 mA 6 k – (a) (b) E2.2 Given the circuits in Fig. E2.2, find (a) R and VS in the circuit in Fig. E2.2a, and (b) find I and R in the circuit in Fig. E2.2b. – 0.4 mA VS ANSWER: (a) I 0.3 mA, P 3.6 mW; (b) VS = 3.6 V, P 2.16 mW. R P=1.6 mW – ± 12 V ANSWER: (a) R=10 k, VS = 4 V; (b) I=20.8 mA, R=576 . R P=0.25 W + I Figure E2.2 (a) (b) E2.3 The power absorbed by Gx in Fig. E2.3 is 50 mW. Find Gx. 10 V Figure E2.3 ± – ANSWER: Gx=500 S. Gx irwin02_025-100hr.qxd 30-06-2010 13:14 Page 31 SECTION 2.2 KIRCHHOFF’S LAWS The circuits we have considered previously have all contained a single resistor, and we have analyzed them using Ohm’s law. At this point we begin to expand our capabilities to handle more complicated networks that result from an interconnection of two or more of these simple elements. We will assume that the interconnection is performed by electrical conductors (wires) that have zero resistance—that is, perfect conductors. Because the wires have zero resistance, the energy in the circuit is in essence lumped in each element, and we employ the term lumped-parameter circuit to describe the network. To aid us in our discussion, we will define a number of terms that will be employed throughout our analysis. As will be our approach throughout this text, we will use examples to illustrate the concepts and define the appropriate terms. For example, the circuit shown R1 v2(t) i1(t) v1(t) R3 R4 R2 ± – 2 i6(t) R4 R5 Kirchhoff’s Laws i2(t) i3(t) R1 R2 Circuit used to illustrate KCL. v1(t) R3 ±– 2.2 Figure 2.5 1 i1(t) 31 3 i4(t) ±– ± v2(t) – i7(t) i5(t) 4 R5 i8(t) 5 (a) (b) in Fig. 2.5a will be used to describe the terms node, loop, and branch. A node is simply a point of connection of two or more circuit elements. The reader is cautioned to note that, although one node can be spread out with perfect conductors, it is still only one node. This is illustrated in Fig. 2.5b, where the circuit has been redrawn. Node 5 consists of the entire bottom connector of the circuit. If we start at some point in the circuit and move along perfect conductors in any direction until we encounter a circuit element, the total path we cover represents a single node. Therefore, we can assume that a node is one end of a circuit element together with all the perfect conductors that are attached to it. Examining the circuit, we note that there are numerous paths through it. A loop is simply any closed path through the circuit in which no node is encountered more than once. For example, starting from node 1, one loop would contain the elements R1 , v2 , R4 , and i1 ; another loop would contain R2 , v1 , v2 , R4 , and i1 ; and so on. However, the path R1 , v1 , R5 , v2 , R3 , and i1 is not a loop because we have encountered node 3 twice. Finally, a branch is a portion of a circuit containing only a single element and the nodes at each end of the element. The circuit in Fig. 2.5 contains eight branches. Given the previous definitions, we are now in a position to consider Kirchhoff’s laws, named after German scientist Gustav Robert Kirchhoff. These two laws are quite simple but extremely important. We will not attempt to prove them because the proofs are beyond our current level of understanding. However, we will demonstrate their usefulness and attempt to make the reader proficient in their use. The first law is Kirchhoff’s current law (KCL), which states that the algebraic sum of the currents entering any node is zero. In mathematical form the law appears as N a ij(t) = 0 j=1 2.7 [hint] KCL is an extremely important and useful law. irwin02_025-100hr.qxd 32 30-06-2010 CHAPTER 2 13:14 Page 32 RESISTIVE CIRCUITS where ij(t) is the jth current entering the node through branch j and N is the number of branches connected to the node. To understand the use of this law, consider node 3 shown in Fig. 2.5. Applying Kirchhoff’s current law to this node yields i2(t)-i4(t)+i5(t)-i7(t)=0 We have assumed that the algebraic signs of the currents entering the node are positive and, therefore, that the signs of the currents leaving the node are negative. If we multiply the foregoing equation by –1, we obtain the expression –i2(t)+i4(t)-i5(t)+i7(t)=0 which simply states that the algebraic sum of the currents leaving a node is zero. Alternatively, we can write the equation as i2(t)+i5(t)=i4(t)+i7(t) which states that the sum of the currents entering a node is equal to the sum of the currents leaving the node. Both of these italicized expressions are alternative forms of Kirchhoff’s current law. Once again it must be emphasized that the latter statement means that the sum of the variables that have been defined entering the node is equal to the sum of the variables that have been defined leaving the node, not the actual currents. For example, ij(t) may be defined entering the node, but if its actual value is negative, there will be positive charge leaving the node. Note carefully that Kirchhoff’s current law states that the algebraic sum of the currents either entering or leaving a node must be zero. We now begin to see why we stated in Chapter 1 that it is critically important to specify both the magnitude and the direction of a current. Recall that current is charge in motion. Based on our background in physics, charges cannot be stored at a node. In other words, if we have a number of charges entering a node, then an equal number must be leaving that same node. Kirchhoff’s current law is based on this principle of conservation of charge. Finally, it is possible to generalize Kirchhoff’s current law to include a closed surface. By a closed surface we mean some set of elements completely contained within the surface that are interconnected. Since the current entering each element within the surface is equal to that leaving the element (i.e., the element stores no net charge), it follows that the current entering an interconnection of elements is equal to that leaving the interconnection. Therefore, Kirchhoff’s current law can also be stated as follows: The algebraic sum of the currents entering any closed surface is zero. EXAMPLE 2.5 SOLUTION Let us write KCL for every node in the network in Fig. 2.5, assuming that the currents leaving the node are positive. The KCL equations for nodes 1 through 5 are -i1(t) + i2(t) + i3(t) = 0 i1(t) - i4(t) + i6(t) = 0 -i2(t) + i4(t) - i5(t) + i7(t) = 0 -i3(t) + i5(t) - i8(t) = 0 -i6(t) - i7(t) + i8(t) = 0 Note carefully that if we add the first four equations, we obtain the fifth equation. What does this tell us? Recall that this means that this set of equations is not linearly independent. We can show that the first four equations are, however, linearly independent. Store this idea in memory because it will become very important when we learn how to write the equations necessary to solve for all the currents and voltages in a network in the following chapter. irwin02_025-100hr.qxd 30-06-2010 13:14 Page 33 SECTION 2.2 KIRCHHOFF’S LAWS 33 EXAMPLE The network in Fig. 2.5 is represented by the topological diagram shown in Fig. 2.6. We wish to find the unknown currents in the network. 2.6 1 I1 60 mA I4 2 I5 3 I6 20 mA 40 mA 4 30 mA Figure 2.6 Topological diagram for the circuit in Fig. 2.5. 5 Assuming the currents leaving the node are positive, the KCL equations for nodes 1 through 4 are SOLUTION -I1 + 0.06 + 0.02 = 0 I1 - I4 + I6 = 0 -0.06 + I4 - I5 + 0.04 = 0 -0.02 + I5 - 0.03 = 0 The first equation yields I1 and the last equation yields I5 . Knowing I5, we can immediately obtain I4 from the third equation. Then the values of I1 and I4 yield the value of I6 from the second equation. The results are I1 = 80 mA, I4 = 70 mA, I5 = 50 mA, and I6 = -10 mA. As indicated earlier, dependent or controlled sources are very important because we encounter them when analyzing circuits containing active elements such as transistors. The following example presents a circuit containing a current-controlled current source. Let us write the KCL equations for the circuit shown in Fig. 2.7. EXAMPLE 2.7 R1 i1(t) 1 R2 i2(t) 2 3 50i2(t) ± v1(t) – i5(t) 4 R3 R4 i3(t) i4(t) Figure 2.7 Circuit containing a dependent current source. irwin02_025-100hr.qxd 34 30-06-2010 CHAPTER 2 13:14 Page 34 RESISTIVE CIRCUITS SOLUTION The KCL equations for nodes 1 through 4 follow: i1(t) + i2(t) - i5(t) = 0 -i2(t) + i3(t) - 50i2(t) = 0 -i1(t) + 50i2(t) + i4(t) = 0 i5(t) - i3(t) - i4(t) = 0 If we added the first three equations, we would obtain the negative of the fourth. What does this tell us about the set of equations? Kirchhoff’s second law, called Kirchhoff’s voltage law (KVL), states that the algebraic sum of the voltages around any loop is zero. As was the case with Kirchhoff’s current law, we will defer the proof of this law and concentrate on understanding how to apply it. Once again the reader is cautioned to remember that we are dealing only with lumped-parameter circuits. These circuits are conservative, meaning that the work required to move a unit charge around any loop is zero. In Chapter 1, we related voltage to the difference in energy levels within a circuit and talked about the energy conversion process in a flashlight. Because of this relationship between voltage and energy, Kirchhoff’s voltage law is based on the conservation of energy. Recall that in Kirchhoff’s current law, the algebraic sign was required to keep track of whether the currents were entering or leaving a node. In Kirchhoff’s voltage law, the algebraic sign is used to keep track of the voltage polarity. In other words, as we traverse the circuit, it is necessary to EXAMPLE 2.8 SOLUTION Let us find I4 and I1 in the network represented by the topological diagram in Fig. 2.6. This diagram is redrawn in Fig. 2.8; node 1 is enclosed in surface 1, and nodes 3 and 4 are enclosed in surface 2. A quick review of the previous example indicates that we derived a value for I4 from the value of I5 . However, I5 is now completely enclosed in surface 2. If we apply KCL to surface 2, assuming the currents out of the surface are positive, we obtain I4 - 0.06 - 0.02 - 0.03 + 0.04 = 0 or I4 = 70 mA which we obtained without any knowledge of I5 . Likewise for surface 1, what goes in must come out and, therefore, I1 = 80 mA. The reader is encouraged to cut the network in Fig. 2.6 into two pieces in any fashion and show that KCL is always satisfied at the boundaries. Figure 2.8 Diagram used to demonstrate KCL for a surface. Surface 1 I1 60 mA I4 I6 20 mA Surface 2 40 mA 30 mA irwin02_025-100hr.qxd 30-06-2010 13:14 Page 35 SECTION 2.2 KIRCHHOFF’S LAWS 35 Learning Assessments E2.4 Given the networks in Fig. E2.3, find (a) I1 in Fig. E2.4a and (b) IT in Fig. E2.4b. ANSWER: (a) I1 = -50 mA; (b) IT = 70 mA. 50 mA ± – IT 10 mA I1 (a) 40 mA 20 mA (b) Figure E2.4 E2.5 Find (a) I1 in the network in Fig. E2.5a and (b) I1 and I2 in the circuit in Fig. E2.5b. I1 12 mA ± – 10 mA I1 4 mA ANSWER: (a) I1 = 6 mA; (b) I1 = 8 mA and I2 = 5 mA. 3 mA I2 4 mA (b) (a) Figure E2.5 E2.6 Find the current ix in the circuits in Fig. E2.6. R 44 mA ix (a) 10ix R1 ANSWER: (a) ix=4 mA; (b) ix=12 mA. 10ix 120 mA ix R2 12 mA (b) Figure E2.6 sum to zero the increases and decreases in energy level. Therefore, it is important we keep track of whether the energy level is increasing or decreasing as we go through each element. In applying KVL, we must traverse any loop in the circuit and sum to zero the increases and decreases in energy level. At this point, we have a decision to make. Do we want to consider a decrease in energy level as positive or negative? We will adopt a policy of considering a decrease in energy level as positive and an increase in energy level as negative. As we move around a loop, we encounter the plus sign first for a decrease in energy level and a negative sign first for an increase in energy level. Finally, we employ the convention Vab to indicate the voltage of point a with respect to point b: that is, the variable for the voltage between point a and point b, with point a considered positive relative to point b. Since the potential is measured between two points, it is convenient to use an arrow between the two points, with the head of the arrow located at the positive node. Note that the double-subscript notation, the ± and – notation, and the singleheaded arrow notation are all the same if the head of the arrow is pointing toward the positive terminal and the first subscript in the double-subscript notation. All of these equivalent forms irwin02_025-100hr.qxd 36 30-06-2010 CHAPTER 2 13:14 Page 36 RESISTIVE CIRCUITS EXAMPLE Consider the circuit shown in Fig. 2.9. If VR1 and VR2 are known quantities, let us find VR3 . 2.9 R1 a + b VR1 R2 VR2 – R3 Figure 2.9 – f Circuit used to illustrate KVL. SOLUTION + 5V ± – 30 V c + – VR+ 3 ±– e d 15 V Starting at point a in the network and traversing it in a clockwise direction, we obtain the equation +VR1 - 5 + VR2 - 15 + VR3 - 30 = 0 which can be written as +VR1 + VR2 + VR3 = 5 + 15 + 30 = 50 Now suppose that VR1 and VR2 are known to be 18 V and 12 V, respectively. Then VR3 = 20 V. EXAMPLE 2.10 Consider the network in Fig. 2.10. a VR1 + VR2 b – + R1 Figure 2.10 Circuit used to explain KVL. 24 V c VR3 + R2 R4 ± – – – d R3 + VR4 – e + 8V ± – 16 V f Let us demonstrate that only two of the three possible loop equations are linearly independent. SOLUTION Note that this network has three closed paths: the left loop, right loop, and outer loop. Applying our policy for writing KVL equations and traversing the left loop starting at point a, we obtain VR1 + VR4 - 16 - 24 = 0 The corresponding equation for the right loop starting at point b is VR2 + VR3 + 8 + 16 - VR4 = 0 The equation for the outer loop starting at point a is VR1 + VR2 + VR3 + 8 - 24 = 0 Note that if we add the first two equations, we obtain the third equation. Therefore, as we indicated in Example 2.5, the three equations are not linearly independent. Once again, we will address this issue in the next chapter and demonstrate that we need only the first two equations to solve for the voltages in the circuit. irwin02_025-100hr.qxd 30-06-2010 13:14 Page 37 SECTION 2.2 KIRCHHOFF’S LAWS 37 for labeling voltages are shown in Fig. 2.11. The usefulness of the arrow notation stems from the fact that we may want to label the voltage between two points that are far apart in a network. In this case, the other notations are often confusing. Figure 2.11 Equivalent forms for labeling voltage. a + a + + + + Vx=Vab 1 Vx=Vo Vo 1 Vx=Vo 1 + Vo Vx=Vab=Vo Vo – – – b b (c) (b) (d) Consider the network in Fig. 2.12a. Let us apply KVL to determine the voltage between two points. Specifically, in terms of the double-subscript notation, let us find Vae and Vec . 16 V + – b 12 V c + R1 24 V ± – R2 R4 f – Vo – – (a) a 1 – + 4V – 24 V 16 V + – ± – – b 12 V + Vec Vae R3 e + 10 V a d f – + 10 V (a) – + 6V 2.11 c + 4V – e + 6V EXAMPLE d (b) The circuit is redrawn in Fig. 2.12b. Since points a and e as well as e and c are not physically close, the arrow notation is very useful. Our approach to determining the unknown voltage is to apply KVL with the unknown voltage in the closed path. Therefore, to determine Vae we can use the path aefa or abcdea. The equations for the two paths in which Vae is the only unknown are Figure 2.12 Network used in Example 2.11. SOLUTION Vae + 10 - 24 = 0 and 16 - 12 + 4 + 6 - Vae = 0 Note that both equations yield Vae = 14 V. Even before calculating Vae , we could calculate Vec using the path cdec or cefabc. However, since Vae is now known, we can also use the path ceabc. KVL for each of these paths is 4 + 6 + Vec = 0 -Vec + 10 - 24 + 16 - 12 = 0 and -Vec - Vae + 16 - 12 = 0 Each of these equations yields Vec = -10 V. [hint] In general, the mathematical representation of Kirchhoff’s voltage law is N a vj(t) = 0 2.8 j=1 where vj(t) is the voltage across the jth branch (with the proper reference direction) in a loop containing N voltages. This expression is analogous to Eq. (2.7) for Kirchhoff’s current law. KVL is an extremely important and useful law. irwin02_025-100hr.qxd 38 30-06-2010 CHAPTER 2 13:14 Page 38 RESISTIVE CIRCUITS Given the network in Fig. 2.13 containing a dependent source, let us write the KVL equations for the two closed paths abda and bcdb. EXAMPLE 2.12 a V + R1 – R1 + VR2 R2 VS ± – Figure 2.13 b 20 VR1 – Network containing a dependent source. c ±– + R3 VR 3 – d SOLUTION The two KVL equations are VR1 + VR2 - VS = 0 20VR1 + VR3 - VR2 = 0 Learning Assessments E2.7 Find IX and I1 in Fig. E2.7. ANSWER: IX = 2 mA, I1 = 4 mA. I1 1.5Ix 6 mA Ix 1 mA Figure E2.7 E2.8 Find Vad and Veb in the network in Fig. E2.8. a + 6V – Figure E2.8 f ±– b c – 4 V+ 24 V ± – +8 V – ANSWER: Vad = 26 V, Veb = 10 V. +12 V – e 6V d E2.9 Find Vbd in the circuit in Fig. E2.9. a+ VR1 ANSWER: Vbd = 11 V. VR2 = 1 V c – – b+ + 12 V Figure E2.9 ± – Vbd – d ± – 10 VR 1 irwin02_025-100hr.qxd 30-06-2010 13:14 Page 39 SECTION 2.3 SINGLE-LOOP CIRCUITS Before proceeding with the analysis of simple circuits, it is extremely important that we emphasize a subtle but very critical point. Ohm’s law as defined by the equation V=IR refers to the relationship between the voltage and current as defined in Fig. 2.14a. If the direction of either the current or the voltage, but not both, is reversed, the relationship between the current and the voltage would be V=–IR. In a similar manner, given the circuit in Fig. 2.14b, if the polarity of the voltage between the terminals A and B is specified as shown, then the direction of the current I is from point B through R to point A. Likewise, in Fig. 2.14c, if the direction of the current is specified as shown, then the polarity of the voltage must be such that point D is at a higher potential than point C and, therefore, the arrow representing the voltage V is from point C to point D. A R A I + I R R C + B B (a) The subtleties associated with Ohm’s law, as described here, are important and must be adhered to in order to ensure that the variables have the proper sign. Circuits used to explain Ohm’s law. V – [hint] Figure 2.14 - V V D I (b) (c) VO LTA G E D I V I S I O N At this point we can begin to apply the laws presented earlier to the analysis of simple circuits. To begin, we examine what is perhaps the simplest circuit—a single closed path, or loop, of elements. Applying KCL to every node in a single-loop circuit reveals that the same current flows through all elements. We say that these elements are connected in series because they carry the same current. We will apply Kirchhoff’s voltage law and Ohm’s law to the circuit to determine various quantities in the circuit. Our approach will be to begin with a simple circuit and then generalize the analysis to more complicated ones. The circuit shown in Fig. 2.15 will serve as a basis for discussion. This circuit consists of an independent voltage source that is in series with two resistors. We have assumed that the current flows in a clockwise direction. If this assumption is correct, the solution of the equations that yields the current will produce a positive value. If the current is actually flowing in the opposite direction, the value of the current variable will simply be negative, indicating that the current is flowing in a direction opposite to that assumed. We have also made voltage polarity assignments for vR1 and vR2 . These assignments have been made using the convention employed in our discussion of Ohm’s law and our choice for the direction of i(t)—that is, the convention shown in Fig. 2.14a. Applying Kirchhoff’s voltage law to this circuit yields 2.3 Single-Loop Circuits i(t) + R1 vR1 – v(t) ± – + R2 vR2 – -v(t) + vR1 + vR2 = 0 or Figure 2.15 v(t) = vR1 + vR2 Single-loop circuit. However, from Ohm’s law we know that vR1 = R1i(t) vR2 = R2i(t) Therefore, v(t)=R1i(t)+R2i(t) Solving the equation for i(t) yields i(t) = 39 v(t) R1 + R2 2.9 irwin02_025-100hr.qxd 40 30-06-2010 CHAPTER 2 13:14 Page 40 RESISTIVE CIRCUITS Knowing the current, we can now apply Ohm’s law to determine the voltage across each resistor: vR1 = R1i(t) = R1 c [hint] v(t) R1 + R2 d = R1 v(t) R1 + R2 vR2 = R2 v(t) R1 + R2 2.10 Similarly, The manner in which voltage divides between two series resistors. 2.11 Though simple, Eqs. (2.10) and (2.11) are very important because they describe the operation of what is called a voltage divider. In other words, the source voltage v(t) is divided between the resistors R1 and R2 in direct proportion to their resistances. In essence, if we are interested in the voltage across the resistor R1 , we bypass the calculation of the current i(t) and simply multiply the input voltage v(t) by the ratio R1 R1 + R2 As illustrated in Eq. (2.10), we are using the current in the calculation, but not explicitly. Note that the equations satisfy Kirchhoff’s voltage law, since -v(t) + EXAMPLE 2.13 R1 R2 v(t) + v(t) = 0 R1 + R2 R1 + R2 Consider the circuit shown in Fig. 2.16. The circuit is identical to Fig. 2.15 except that R1 is a variable resistor such as the volume control for a radio or television set. Suppose that VS = 9 V, R1 = 90 k, and R2 = 30 k. I R1 VS ± – + R2 Figure 2.16 V2 – Voltage-divider circuit. Let us examine the change in both the voltage across R2 and the power absorbed in this resistor as R1 is changed from 90 k to 15 k. SOLUTION Since this is a voltage-divider circuit, the voltage V2 can be obtained directly as V2 = c = c R2 d VS R1 + R2 30k d (9) 90k + 30k = 2.25 V irwin02_025-100hr.qxd 30-06-2010 13:14 Page 41 SECTION 2.3 SINGLE-LOOP CIRCUITS 41 Now suppose that the variable resistor is changed from 90 k to 15 k. Then V2 = c 30k d9 30k + 15k = 6V The direct voltage-divider calculation is equivalent to determining the current I and then using Ohm’s law to find V2 . Note that the larger voltage is across the larger resistance. This voltage-divider concept and the simple circuit we have employed to describe it are very useful because, as will be shown later, more complicated circuits can be reduced to this form. Finally, let us determine the instantaneous power absorbed by the resistor R2 under the two conditions R1 = 90 k and R1 = 15 k. For the case R1 = 90 k, the power absorbed by R2 is 2 9 b (30k) 120k = 0.169 mW P2 = I 2R2 = a In the second case 9 2 b (30k) 45k = 1.2 mW P2 = a The current in the first case is 75 A, and in the second case it is 200 A. Since the power absorbed is a function of the square of the current, the power absorbed in the two cases is quite different. Let us now demonstrate the practical utility of this simple voltage-divider network. Consider the circuit in Fig. 2.17a, which is an approximation of a high-voltage dc transmission facility. We have assumed that the bottom portion of the transmission line is a perfect conductor and will justify this assumption in the next chapter. The load can be represented by a resistor of value 183.5 . Therefore, the equivalent circuit of this network is shown in Fig. 2.17b. Line resistance is 0.04125 /mile 16.5 400 kV 2.14 2 kA 2 kA ± – EXAMPLE ± – Load + Vload 400 kV Perfect conductor 183.5 – Figure 2.17 400-mile transmission line (a) (b) A high-voltage dc transmission facility. Let us determine both the power delivered to the load and the power losses in the line. Using voltage division, the load voltage is Vload = c 183.5 d 400k 183.5 + 16.5 = 367 kV SOLUTION irwin02_025-100hr.qxd 42 30-06-2010 CHAPTER 2 13:14 Page 42 RESISTIVE CIRCUITS The input power is 800 MW and the power transmitted to the load is Pload = I 2Rload = 734 MW Therefore, the power loss in the transmission line is Pline = Pin - Pload = I 2Rline = 66 MW Since P=VI, suppose now that the utility company supplied power at 200 kV and 4 kA. What effect would this have on our transmission network? Without making a single calculation, we know that because power is proportional to the square of the current, there would be a large increase in the power loss in the line and, therefore, the efficiency of the facility would decrease substantially. That is why, in general, we transmit power at high voltage and low current. M U LT I P L E -S O U R C E / R E S I S TO R N E T W O R KS At this point we wish to extend our analysis to include a multiplicity of voltage sources and resistors. For example, consider the circuit shown in Fig. 2.18a. Here we have assumed that the current flows in a clockwise direction, and we have defined the variable i(t) accordingly. This may or may not be the case, depending on the value of the various voltage sources. Kirchhoff’s voltage law for this circuit is +vR1 + v2(t) - v3(t) + vR2 + v4(t) + v5(t) - v1(t) = 0 or, using Ohm’s law, AR1 + R2 Bi(t) = v1(t) - v2(t) + v3(t) - v4(t) - v5(t) which can be written as where AR1 + R2 Bi(t) = v(t) v(t) = v1(t) + v3(t) - C v2(t) + v4(t) + v5(t)D so that under the preceding definitions, Fig. 2.18a is equivalent to Fig. 2.18b. In other words, the sum of several voltage sources in series can be replaced by one source whose value is the algebraic sum of the individual sources. This analysis can, of course, be generalized to a circuit with N series sources. Figure 2.18 v2(t) v i(t)+ R1 – Equivalent circuits with multiple sources. ±– R1 v1(t) ± – - v3(t) + i(t) R1 + R2 vR2 v5(t) + – v(t) ± – R2 + v4(t) (a) (b) irwin02_025-100hr.qxd 30-06-2010 13:14 Page 43 SECTION 2.3 + vR1 - + R1 vR2 - + R2 vR3 43 - R3 i(t) v(t) SINGLE-LOOP CIRCUITS + R4 vR4 i(t) ± – + R5 RN v(t) ± – vR5 - RS=R1+R2+R3+…+RN -v + RN (a) (b) Figure 2.19 Now consider the circuit with N resistors in series, as shown in Fig. 2.19a. Applying Kirchhoff’s voltage law to this circuit yields Equivalent circuits. v(t) = vR1 + vR2 + p + vRN = R1i(t) + R2i(t) + p + RNi(t) and therefore, v(t)=RSi(t) 2.12 where RS = R1 + R2 + p + RN 2.13 and hence, i(t) = v(t) 2.14 RS Note also that for any resistor Ri in the circuit, the voltage across Ri is given by the expression vRi = Ri v(t) RS 2.15 which is the voltage-division property for multiple resistors in series. Equation (2.13) illustrates that the equivalent resistance of N resistors in series is simply the sum of the individual resistances. Thus, using Eq. (2.13), we can draw the circuit in Fig. 2.19b as an equivalent circuit for the one in Fig. 2.19a. Given the circuit in Fig. 2.20a, let us find I, Vbd , and the power absorbed by the 30-k resistor. Finally, let us use voltage division to find Vbc . ± – I 10 k b 20 k ± – 6V 40 k c 12 V ± – a EXAMPLE 2.15 b 6V 20 k 30 k e c d (a) (b) Figure 2.20 Circuit used in Example 2.15. irwin02_025-100hr.qxd 44 30-06-2010 CHAPTER 2 13:14 Page 44 RESISTIVE CIRCUITS SOLUTION KVL for the network yields the equation 10kI + 20kI + 12 + 30kI - 6 = 0 60kI = -6 I = -0.1 mA Therefore, the magnitude of the current is 0.1 mA, but its direction is opposite to that assumed. The voltage Vbd can be calculated using either of the closed paths abdea or bcdb. The equations for both cases are 10kI + Vbd + 30kI - 6 = 0 and 20kI + 12 - Vbd = 0 Using I=–0.1 mA in either equation yields Vbd = 10 V. Finally, the power absorbed by the 30-k resistor is P = I 2R = 0.3 mW Now from the standpoint of determining the voltage Vbc , we can simply add the sources since they are in series, add the remaining resistors since they are in series, and reduce the network to that shown in Fig. 2.20b. Then Vbc = 20k (-6) 20k + 40k = -2 V EXAMPLE 2.16 A dc transmission facility is modeled by the approximate circuit shown in Fig. 2.21. If the load voltage is known to be Vload = 458.3 kV, we wish to find the voltage at the sending end of the line and the power loss in the line. IL Rline 20 Figure 2.21 ± VS – + Rload 220 Circuit used in Example 2.16. SOLUTION Vload=458.3 kV - Knowing the load voltage and load resistance, we can obtain the line current using Ohm’s law: IL = 458.3k兾220 = 2.083 kA The voltage drop across the line is Vline = (IL)(Rline) = 41.66 kV irwin02_025-100hr.qxd 30-06-2010 13:14 Page 45 SECTION 2.3 SINGLE-LOOP CIRCUITS 45 Now, using KVL, VS = Vline + Vload = 500 kV Note that since the network is simply a voltage-divider circuit, we could obtain VS immediately from our knowledge of Rline, Rload, and Vload. That is, Vload = c Rload d VS Rload + Rline and VS is the only unknown in this equation. The power absorbed by the line is Pline = I 2LRline = 86.79 MW Problem-Solving Strategy Step 1. Define a current i(t). We know from KCL that there is only one current for a Single-Loop Circuits single-loop circuit. This current is assumed to be flowing either clockwise or counterclockwise around the loop. Step 2. Using Ohm’s law, define a voltage across each resistor in terms of the defined current. Step 3. Apply KVL to the single-loop circuit. Step 4. Solve the single KVL equation for the current i(t). If i(t) is positive, the current is flowing in the direction assumed; if not, then the current is actually flowing in the opposite direction. Learning Assessments E2.10 Find I and Vbd in the circuit in Fig. E2.10. a b 12 V c ±– I 80 k ± – ANSWER: I 0.05 mA and Vbd = 10 V. 6V 40 k d Figure E2.10 ANSWER: VS = 9 V. E2.11 In the network in Fig. E2.11, if Vad is 3 V, find VS . a ±– b 25 k c VS 20 k Figure E2.11 15 k d irwin02_025-100hr.qxd 46 30-06-2010 CHAPTER 2 13:14 Page 46 RESISTIVE CIRCUITS 2.4 Single-Node-Pair Circuits CURRENT DIVISION An important circuit is the single-node-pair circuit. If we apply KVL to every loop in a single-node-pair circuit, we discover that all of the elements have the same voltage across them and, therefore, are said to be connected in parallel. We will, however, apply Kirchhoff’s current law and Ohm’s law to determine various unknown quantities in the circuit. Following our approach with the single-loop circuit, we will begin with the simplest case and then generalize our analysis. Consider the circuit shown in Fig. 2.22. Here we have an independent current source in parallel with two resistors. + Figure 2.22 Simple parallel circuit. i(t) R1 R2 v(t) i1(t) i2(t) – Since all of the circuit elements are in parallel, the voltage v(t) appears across each of them. Furthermore, an examination of the circuit indicates that the current i(t) is into the upper node of the circuit and the currents i1(t) and i2(t) are out of the node. Since KCL essentially states that what goes in must come out, the question we must answer is how i1(t) and i2(t) divide the input current i(t). Applying Kirchhoff’s current law to the upper node, we obtain i(t)=i1(t)+i2(t) and, employing Ohm’s law, we have i(t) = v(t) v(t) + R1 R2 1 1 = a + b v(t) R1 R2 v(t) = Rp [hint] The parallel resistance equation. where 1 1 1 = + Rp R1 R2 Rp = R1 R2 R1 + R2 2.16 2.17 Therefore, the equivalent resistance of two resistors connected in parallel is equal to the product of their resistances divided by their sum. Note also that this equivalent resistance Rp is always less than either R1 or R2 . Hence, by connecting resistors in parallel we reduce the overall resistance. In the special case when R1 = R2 , the equivalent resistance is equal to half of the value of the individual resistors. The manner in which the current i(t) from the source divides between the two branches is called current division and can be found from the preceding expressions. For example, v(t) = Rp i(t) = R1 R2 i(t) R1 + R2 2.18 irwin02_025-100hr.qxd 30-06-2010 13:14 Page 47 SECTION 2.4 S I N G L E - N O D E - PA I R C I R C U I T S and i1(t) = [hint] v(t) R1 R2 i1(t) = i(t) R1 + R2 47 2.19 The manner in which current divides between two parallel resistors. and i2(t) = v(t) R2 R1 = i(t) R1 + R2 2.20 Eqs. (2.19) and (2.20) are mathematical statements of the current-division rule. Given the network in Fig. 2.23a, let us find I1 , I2 , and Vo . First, it is important to recognize that the current source feeds two parallel paths. To emphasize this point, the circuit is redrawn as shown in Fig. 2.23b. Applying current division, we obtain I1 = c 40k + 80k d A0.9 * 10-3 B 60k + (40k + 80k) = 0.6 mA and I2 = c 60k d A0.9 * 10-3 B 60k + (40k + 80k) = 0.3 mA Note that the larger current flows through the smaller resistor, and vice versa. In addition, note that if the resistances of the two paths are equal, the current will divide equally between them. KCL is satisfied since I1+I2=0.9 mA. The voltage Vo can be derived using Ohm’s law as Vo = 80kI2 = 24 V The problem can also be approached in the following manner. The total resistance seen by the current source is 40 k, that is, 60 k in parallel with the series combination of 40 k and 80 k, as shown in Fig. 2.23c. The voltage across the current source is then V1 = A0.9 * 10-3 B 40k = 36 V Now that V1 is known, we can apply voltage division to find Vo: Vo = a = a 80k bV 80k + 40k 1 80k b 36 120k = 24 V EXAMPLE SOLUTION 2.17 irwin02_025-100hr.qxd 48 30-06-2010 CHAPTER 2 13:14 Page 48 RESISTIVE CIRCUITS I2 I2 + V1 0.9 mA 40 k 60 k I1 0.9 mA 60 k – Vo 80 k – – + + Vo 80 k 40 k V1 + I1 + 0.9 mA – (a) V1 40 k – (b) (c) Figure 2.23 Circuits used in Example 2.17. EXAMPLE A typical car stereo consists of a 2-W audio amplifier and two speakers represented by the diagram shown in Fig. 2.24a. The output circuit of the audio amplifier is in essence a 430-mA current source, and each speaker has a resistance of 4 . Let us determine the power absorbed by the speakers. 2.18 SOLUTION The audio system can be modeled as shown in Fig. 2.24b. Since the speakers are both 4- devices, the current will split evenly between them, and the power absorbed by each speaker is P = I 2R = A215 * 10-3 B (4) 2 = 184.9 mW Figure 2.24 Circuits used in Example 2.18. Audio amplifier 4 430 mA (a) 4 (b) Learning Assessment E2.12 Find the currents I1 and I2 and the power absorbed by the 40-k resistor in the network in Fig. E2.12. 16 mA 40 k Figure E2.12 120 k I1 I2 ANSWER: I1 = 12 mA, I2 = -4 mA, and P40 k = 5.76 W. irwin02_025-100hr.qxd 30-06-2010 13:14 Page 49 SECTION 2.4 i2(t) + i1(t) v(t) S I N G L E - N O D E - PA I R C I R C U I T S i5(t) R1 i3(t) i4(t) + R2 i6(t) io(t) R1 R2 v(t) – – (b) (a) Figure 2.25 Equivalent circuits. M U LT I P L E -S O U R C E / R E S I S TO R N E T W O R KS Let us now extend our analysis to include a multiplicity of current sources and resistors in parallel. For example, consider the circuit shown in Fig. 2.25a. We have assumed that the upper node is v(t) volts positive with respect to the lower node. Applying Kirchhoff’s current law to the upper node yields i1(t)-i2(t)-i3(t)+i4(t)-i5(t)-i6(t)=0 or i1(t)-i3(t)+i4(t)-i6(t)=i2(t)+i5(t) The terms on the left side of the equation all represent sources that can be combined algebraically into a single source; that is, io(t)=i1(t)-i3(t)+i4(t)-i6(t) which effectively reduces the circuit in Fig. 2.25a to that in Fig. 2.25b. We could, of course, generalize this analysis to a circuit with N current sources. Using Ohm’s law, we can express the currents on the right side of the equation in terms of the voltage and individual resistances so that the KCL equation reduces to io(t) = a 1 1 + b v(t) R1 R2 Now consider the circuit with N resistors in parallel, as shown in Fig. 2.26a. Applying Kirchhoff’s current law to the upper node yields + io(t) v(t) i1(t) i2(t) iN(t) R1 R2 RN Figure 2.26 + io(t) v(t) – – (b) (a) io(t) = i1(t) + i2(t) + p + iN(t) = a 1 1 1 + + p + b v(t) R1 R2 RN 2.21 or io(t) = v(t) 2.22 Rp where N 1 1 = a Rp i = 1 Ri 2.23 Equivalent circuits. Rp 49 irwin02_025-100hr.qxd 50 30-06-2010 CHAPTER 2 13:14 Page 50 RESISTIVE CIRCUITS so that as far as the source is concerned, Fig. 2.26a can be reduced to an equivalent circuit, as shown in Fig. 2.26b. The current division for any branch can be calculated using Ohm’s law and the preceding equations. For example, for the jth branch in the network of Fig. 2.26a, ij(t) = v(t) Rj Using Eq. (2.22), we obtain ij(t) = Rp Rj io(t) 2.24 which defines the current-division rule for the general case. EXAMPLE Given the circuit in Fig. 2.27a, we wish to find the current in the 12-k load resistor. 2.19 SOLUTION To simplify the network in Fig. 2.27a, we add the current sources algebraically and combine the parallel resistors in the following manner: 1 1 1 1 = + + Rp 18k 9k 12k Rp = 4 k Using these values we can reduce the circuit in Fig. 2.27a to that in Fig. 2.27b. Now, applying current division, we obtain IL = - c 4k d A1 * 10-3 B 4k + 12k = -0.25 mA 1 mA 18 k 2 mA 9 k 12 k 4 mA (a) IL IL RL=12 k 4 k 12 k 1 mA (b) Figure 2.27 Circuits used in Example 2.19. Problem-Solving Strategy Single-Node-Pair Circuits Step 1. Define a voltage v(t) between the two nodes in this circuit. We know from KVL that there is only one voltage for a single-node-pair circuit. A polarity is assigned to the voltage such that one of the nodes is assumed to be at a higher potential than the other node, which we will call the reference node. Step 2. Using Ohm’s law, define a current flowing through each resistor in terms of the defined voltage. Step 3. Apply KCL at one of the two nodes in the circuit. Step 4. Solve the single KCL equation for v(t). If v(t) is positive, then the reference node is actually at a lower potential than the other node; if not, the reference node is actually at a higher potential than the other node irwin02_025-100hr.qxd 30-06-2010 13:14 Page 51 SECTION 2.5 S E R I E S A N D PA R A L L E L R E S I S T O R C O M B I N AT I O N S 51 Learning Assessment E2.13 Find the power absorbed by the 6-kΩ resistor in the network in Fig. E2.13. ANSWER: P 2.67 mW. 4 mA 6 mA 4 k 6 k 12 k Figure E2.13 2.5 We have shown in our earlier developments that the equivalent resistance of N resistors in series is RS = R1 + R2 + p + RN 2.25 Series and Parallel Resistor Combinations and the equivalent resistance of N resistors in parallel is found from 1 1 1 1 = + + p + Rp R1 R2 RN 2.26 Let us now examine some combinations of these two cases. EXAMPLE We wish to determine the resistance at terminals A-B in the network in Fig. 2.28a. SOLUTION Starting at the opposite end of the network from the terminals and combining resistors as shown in the sequence of circuits in Fig. 2.28, we find that the equivalent resistance at the terminals is 5 k. 2 k 2 k 2 k 10 k A RAB 2 k A 6 k 4 k 6 k 6 k 1 k B RAB 4 k 6 k 6 k 12 k=10 k+ (6 k in parallel with 3 k) B 9 k 9 k 2 k (b) (a) 2 k 2 k A RAB 2.20 A 4 k 6 k=2 k+ (6 k in parallel with 12 k) 6 k B RAB 4 k 12 k=9 k+ (6 k in parallel with 6 k) B 9 k (c) (d) 2 k A RAB 3 k=(4 k in parallel with 12 k) Figure 2.28 B (e) Simplification of a resistance network. irwin02_025-100hr.qxd 52 30-06-2010 CHAPTER 2 13:14 Page 52 RESISTIVE CIRCUITS Learning Assessment E2.14 Find the equivalent resistance at the terminals A-B in the network in Fig. E2.14. 3 k 6 k A ANSWER: RAB 22 k. 18 k 6 k RAB 10 k Figure E2.14 B Problem-Solving Strategy Simplifying Resistor Combinations When trying to determine the equivalent resistance at a pair of terminals of a network composed of an interconnection of numerous resistors, it is recommended that the analysis begin at the end of the network opposite the terminals. Two or more resistors are combined to form a single resistor, thus simplifying the network by reducing the number of components as the analysis continues in a steady progression toward the terminals. The simplification involves the following: Step 1. Resistors in series. Resistors R1 and R2 are in series if they are connected end to end with one common node and carry exactly the same current. They can then be combined into a single resistor RS, where RS=R1+R2 . Step 2. Resistors in parallel. Resistors R1 and R2 are in parallel if they are connected to the same two nodes and have exactly the same voltage across their terminals. They can then be combined into a single resistor Rp , where Rp = R1 R2兾AR1 + R2 B. These two combinations are used repeatedly, as needed, to reduce the network to a single resistor at the pair of terminals. Learning Assessment E2.15 Find the equivalent resistance at the terminals A-B in the circuit in Fig. E2.15. 4 k ANSWER: RAB 3 k. 4 k A RAB Figure E2.15 6 k 3 k 12 k 8 k B E2.16 Find RAB in Fig. E2.16. ANSWER: RAB 12 k. A RAB Figure E2.16 4 k 8 k B 4 k 3 k 12 k 4 k 2 k 6 k 2 k 2 k irwin02_025-100hr.qxd 30-06-2010 13:14 Page 53 SECTION 2.5 S E R I E S A N D PA R A L L E L R E S I S T O R C O M B I N AT I O N S A standard dc current-limiting power supply shown in Fig. 2.29a provides 0–18 V at 3 A to a load. The voltage drop, VR , across a resistor, R, is used as a current-sensing device, fed back to the power supply and used to limit the current I. That is, if the load is adjusted so that the current tries to exceed 3 A, the power supply will act to limit the current to that value. The feedback voltage, VR , should typically not exceed 600 mV. If we have a box of standard 0.1- Æ , 5-W resistors, let us determine the configuration of these resistors that will provide VR = 600 mV when the current is 3 A. dc power supply 2.21 0.1 0A + EXAMPLE I + R VR R - R 0A Load 0.1 All resistors 0.1 (a) (b) (c) Figure 2.29 Circuits used in Example 2.21. Using Ohm’s law, the value of R should be SOLUTION R = = VR I 0.6 3 = 0.2 Therefore, two 0.1- resistors connected in series, as shown in Fig. 2.29b, will provide the proper feedback voltage. Suppose, however, that the power supply current is to be limited to 9 A. The resistance required in this case to produce VR = 600 mV is R = 0.6 9 = 0.0667 We must now determine how to interconnect the 0.1- Æ resistor to obtain R = 0.0667 . Since the desired resistance is less than the components available (i.e., 0.1- Æ ), we must connect the resistors in some type of parallel configuration. Since all the resistors are of equal value, note that three of them connected in parallel would provide a resistance of one-third their value, or 0.0333 . Then two such combinations connected in series, as shown in Fig. 2.29c, would produce the proper resistance. 53 irwin02_025-100hr.qxd 54 30-06-2010 CHAPTER 2 13:14 Page 54 RESISTIVE CIRCUITS Finally, we must check to ensure that the configurations in Figs. 2.29b and c have not exceeded the power rating of the resistors. In the first case, the current I = 3 A is present in each of the two series resistors. Therefore, the power absorbed in each resistor is P = I 2R = (3)2(0.1) = 0.9 W which is well within the 5-W rating of the resistors. In the second case, the current I = 9 A. The resistor configuration for R in this case is a series combination of two sets of three parallel resistors of equal value. Using current division, we know that the current I will split equally among the three parallel paths and, hence, the current in each resistor will be 3 A. Therefore, once again, the power absorbed by each resistor is within its power rating. R E S I S TO R S P EC I F I C AT I O N S Some important parameters that are used to specify resistors are the resistor’s value, tolerance, and power rating. The tolerance specifications for resistors are typically 5% and 10%. A listing of standard resistor values with their specified tolerances is shown in Table 2.1. The power rating for a resistor specifies the maximum power that can be dissipated by the resistor. Some typical power ratings for resistors are 1兾4 W, 1兾2 W, 1 W, 2 W, and so forth, up to very high values for high-power applications. Thus, in selecting a resistor for some particular application, one important selection criterion is the expected power dissipation. TABLE 2.1 Standard resistor values for 5% and 10% tolerances (values available with a 10% tolerance shown in boldface) 1.0 1.1 1.2 1.3 1.5 1.6 1.8 2.0 2.2 2.4 2.7 3.0 3.3 3.6 3.9 4.3 4.7 5.1 5.6 6.2 6.8 7.5 8.2 9.1 10 11 12 13 15 16 18 20 22 24 27 30 33 36 39 43 47 51 56 62 68 75 82 91 100 110 120 130 150 160 180 200 220 240 270 300 330 360 390 430 470 510 560 620 680 750 820 910 1.0k 1.1k 1.2k 1.3k 1.5k 1.6k 1.8k 2.0k 2.2k 2.4k 2.7k 3.0k 3.3k 3.6k 3.9k 4.3k 4.7k 5.1k 5.6k 6.2k 6.8k 7.5k 8.2k 9.1k 10k 11k 12k 13k 15k 16k 18k 20k 22k 24k 27k 30k 33k 36k 39k 43k 47k 51k 56k 62k 68k 75k 82k 91k 100k 110k 120k 130k 150k 160k 180k 200k 220k 240k 270k 300k 330k 360k 390k 430k 470k 510k 560k 620k 680k 750k 820k 910k 1.0M 1.1M 1.2M 1.3M 1.5M 1.6M 1.8M 2.0M 2.2M 2.4M 2.7M 3.0M 3.3M 3.6M 3.9M 4.3M 4.7M 5.1M 5.6M 6.2M 6.8M 7.5M 8.2M 9.1M 10M 11M 12M 13M 15M 16M 18M 20M 22M irwin02_025-100hr.qxd 30-06-2010 13:14 SECTION 2.6 Page 55 55 C I R C U I T S W I T H S E R I E S - PA R A L L E L C O M B I N AT I O N S O F R E S I S T O R S Given the network in Fig. 2.30, we wish to find the range for both the current and power dissipation in the resistor if R is a 2.7-k resistor with a tolerance of 10%. Using the equations I=V/R=10/R and P = V2兾R = 100兾R, the minimum and maximum values for the resistor, current, and power are outlined next. Minimum resistor value=R(1-0.1)=0.9 R=2.43 k Maximum resistor value=R(1+0.1)=1.1 R=2.97 k Minimum current value=10/2970=3.37 mA Maximum current value=10/2430=4.12 mA Minimum power value=100/2970=33.7 mW Maximum power value=100/2430=41.2 mW EXAMPLE 2.22 SOLUTION I 10 V ± – R Figure 2.30 Thus, the ranges for the current and power are 3.37 mA to 4.12 mA and 33.7 mW to 41.2 mW, respectively. Circuit used in Example 2.22. Given the network shown in Fig. 2.31: (a) find the required value for the resistor R; (b) use Table 2.1 to select a standard 10% tolerance resistor for R; (c) using the resistor selected in (b), determine the voltage across the 3.9-k resistor; (d) calculate the percent error in the voltage V1 , if the standard resistor selected in (b) is used; and (e) determine the power rating for this standard component. a. Using KVL, the voltage across R is 19 V. Then using Ohm’s law, the current in the loop is EXAMPLE 2.23 SOLUTION I=5/3.9k=1.282 mA The required value of R is then R R=19/0.001282=14.82 k b. As shown in Table 2.1, the nearest standard 10% tolerance resistor is 15 k. 24 V ± – + c. Using the standard 15-k resistor, the actual current in the circuit is 3.9 k – I=24/18.9k=1.2698 mA and the voltage across the 3.9-k resistor is V=IR=(0.0012698)(3.9k)=4.952 V V1=5 V Figure 2.31 Circuit used in Example 2.23. d. The percent error involved in using the standard resistor is % Error=(4.952-5)/5*100=–0.96% e. The power absorbed by the resistor R is then P = IR = (0.0012698)2(15k) = 24.2 mW Therefore, even a quarter-watt resistor is adequate in this application. At this point we have learned many techniques that are fundamental to circuit analysis. Now we wish to apply them and show how they can be used in concert to analyze circuits. We will illustrate their application through a number of examples that will be treated in some detail. 2.6 Circuits with Series-Parallel Combinations of Resistors irwin02_025-100hr.qxd 56 30-06-2010 CHAPTER 2 13:14 RESISTIVE CIRCUITS EXAMPLE 2.24 I1 We wish to find all the currents and voltages labeled in the ladder network shown in Fig. 2.32a. I3 I5 I1 I3 9 k 12 V Page 56 ± – I2 9 k 3 k I4 + + Va 6 k Vb 4 k – – I1 I3 9 k + Vc 3 k – 12 V ± – I5 (a) (b) 3 1 —V — mA +2 – 2 1 mA +9 V – I1 9 k 12 V 9 k + ± – Va – 3 k I2 3 k + + Va 6 k Vb 3 k – – 12 V ± – 1 — mA + 2 6 k 3V – (c) 3 k 9 1 —V — mA +8 – 8 3 — mA 8 9 k + 3 — V 4 k 2 – + 3 — V 3 k 8 – (d) Figure 2.32 Analysis of a ladder network. SOLUTION To begin our analysis of the network, we start at the right end of the circuit and combine the resistors to determine the total resistance seen by the 12-V source. This will allow us to calculate the current I1 . Then employing KVL, KCL, Ohm’s law, and/or voltage and current division, we will be able to calculate all currents and voltages in the network. At the right end of the circuit, the 9-k and 3-k resistors are in series and, thus, can be combined into one equivalent 12-k resistor. This resistor is in parallel with the 4-k resistor, and their combination yields an equivalent 3-k resistor, shown at the right edge of the circuit in Fig. 2.32b. In Fig. 2.32b the two 3-k resistors are in series, and their combination is in parallel with the 6-k resistor. Combining all three resistances yields the circuit shown in Fig. 2.32c. Applying Kirchhoff’s voltage law to the circuit in Fig. 2.32c yields I1(9k + 3k) = 12 I1 = 1 mA Va can be calculated from Ohm’s law as Va = I1(3k) = 3V or, using Kirchhoff’s voltage law, Va = 12 - 9kI1 = 12 - 9 = 3V Knowing I1 and Va , we can now determine all currents and voltages in Fig. 2.32b. Since Va = 3 V, the current I2 can be found using Ohm’s law as 3 6k 1 = mA 2 I2 = irwin02_025-100hr.qxd 30-06-2010 13:14 SECTION 2.6 Page 57 C I R C U I T S W I T H S E R I E S - PA R A L L E L C O M B I N AT I O N S O F R E S I S T O R S Then, using Kirchhoff’s current law, we have I1 = I2 + I3 1 * 10-3 + I3 1 * 10-3 = 2 I3 = 1 mA 2 Note that the I3 could also be calculated using Ohm’s law: Va = (3k + 3k)I3 3 I3 = 6k 1 = mA 2 Applying Kirchhoff’s voltage law to the right-hand loop in Fig. 2.32b yields Va - Vb = 3kI3 3 - Vb = Vb = 3 2 3 V 2 or, since Vb is equal to the voltage drop across the 3-k resistor, we could use Ohm’s law as Vb = 3kI3 3 = V 2 We are now in a position to calculate the final unknown currents and voltages in Fig. 2.32a. Knowing Vb , we can calculate I4 using Ohm’s law as Vb = 4kI4 3 2 I4 = 4k 3 = mA 8 Then, from Kirchhoff’s current law, we have I3 = I4 + I5 1 3 * 10-3 = * 10-3 + I5 2 8 1 I5 = mA 8 We could also have calculated I5 using the current-division rule. For example, 4k I 4k + (9k + 3k) 3 1 = mA 8 I5 = 57 irwin02_025-100hr.qxd 58 30-06-2010 CHAPTER 2 13:14 Page 58 RESISTIVE CIRCUITS Finally, Vc can be computed as Vc = I5(3k) 3 = V 8 Vc can also be found using voltage division (i.e., the voltage Vb will be divided between the 9-k and 3-k resistors). Therefore, Vc = c = 3k dV 3k + 9k b 3 V 8 Note that Kirchhoff’s current law is satisfied at every node and Kirchhoff’s voltage law is satisfied around every loop, as shown in Fig. 2.32d. The following example is, in essence, the reverse of the previous example in that we are given the current in some branch in the network and are asked to find the value of the input source. EXAMPLE Given the circuit in Fig. 2.33 and I4 = 1兾2 mA, let us find the source voltage Vo . 2.25 SOLUTION If I4 = 1兾2 mA, then from Ohm’s law, Vb = 3 V. Vb can now be used to calculate I3 = 1 mA. Kirchhoff’s current law applied at node y yields I2 = I3 + I4 = 1.5 mA Then, from Ohm’s law, we have Va = A1.5 * 10-3 B(2k) = 3V Since Va + Vb is now known, I5 can be obtained: I5 = Va + Vb 3k + 1k = 1.5 mA Applying Kirchhoff’s current law at node x yields I1 = I2 + I5 = 3 mA Now KVL applied to any closed path containing Vo will yield the value of this input source. For example, if the path is the outer loop, KVL yields -Vo + 6kI1 + 3kI5 + 1kI5 + 4kI1 = 0 Since I1 = 3 mA and I5 = 1.5 mA, Vo = 36 V irwin02_025-100hr.qxd 30-06-2010 13:14 SECTION 2.6 Page 59 C I R C U I T S W I T H S E R I E S - PA R A L L E L C O M B I N AT I O N S O F R E S I S T O R S 59 If we had selected the path containing the source and the points x, y, and z, we would obtain -Vo + 6kI1 + Va + Vb + 4kI1 = 0 Once again, this equation yields Vo = 36 V 6 k I1 x 3 k Figure 2.33 I5 Example circuit for analysis. + Va 2 k – I2 y Vo ± – 1 k + Vb 3 k I3 6 k I4 – z 4 k Problem-Solving Strategy Step 1. Systematically reduce the resistive network so that the resistance seen by the source is represented by a single resistor. Step 2. Determine the source current for a voltage source or the source voltage if a current source is present. Step 3. Expand the network, retracing the simplification steps, and apply Ohm’s law, KVL, KCL, voltage division, and current division to determine all currents and voltages in the network. Analyzing Circuits Containing a Single Source and a Series-Parallel Interconnection of Resistors Learning Assessments E2.17 Find Vo in the network in Fig. E2.17. 20 k ANSWER: Vo = 2 V. 40 k + 12 V ± – 30 k 20 k Vo – Figure E2.17 E2.18 Find VS in the circuit in Fig. E2.18. ANSWER: VS = 9 V. 20 k VS ± – 60 k 0.1 mA Figure E2.18 120 k irwin02_025-100hr.qxd 60 30-06-2010 CHAPTER 2 13:14 Page 60 RESISTIVE CIRCUITS E2.19 Find IS in the circuit in Fig. E2.19. ANSWER: IS = 0.3 mA. 90 k + IS 60 k 3V 30 k – Figure E2.19 E2.20 Find V1 in Fig. E2.20. ANSWER: V1=12 V. + V1 3 k – 6 k 25 mA 15 mA 6 k Figure E2.20 E2.21 Find I0 in Fig. E2.21. ANSWER: I0=- 4 mA. 4 k 12 k 9 mA 3 k 6 k I0 Figure E2.21 E2.22 Find Vo, V1, and V2 in Fig. E2.22. ANSWER: Vo=3.33 V, V1=- 4 V, V2=4 V. – 10 k 20 k 3 k V1 + Figure E2.22 4 k 4 k + Vo – + 5 k 10 k 16 V + – 12 k 8 k V2 – irwin02_025-100hr.qxd 30-06-2010 13:14 Page 61 SECTION 2.7 61 W Y E ∆ D E LTA T R A N S F O R M AT I O N S E2.23 Find V0 and V1 in Fig. E2.23. ANSWER: V0 = -60 V, V1 = 10 V. 10 k 6 k – 4 k + Vo 12 k V1 20 mA + 4 k – Figure E2.23 2.7 To provide motivation for this topic, consider the circuit in Fig. 2.34. Note that this network has essentially the same number of elements as contained in our recent examples. However, when we attempt to reduce the circuit to an equivalent network containing the source V1 and an equivalent resistor R, we find that nowhere is a resistor in series or parallel with another. Therefore, we cannot attack the problem directly using the techniques that we have learned thus far. We can, however, replace one portion of the network with an equivalent circuit, and this conversion will permit us, with ease, to reduce the combination of resistors to a single equivalent resistance. This conversion is called the wye-to-delta or delta-to-wye transformation. Wye ∆ Delta Transformations a I1 a R1 R2 Ra R3 V1 ± – R4 R1 R5 R2 c R3 R6 Rc b c (a) Figure 2.34 Figure 2.35 Network used to illustrate the need for the wye ∆ delta transformation. Delta and wye resistance networks. Rb Consider the networks shown in Fig. 2.35. Note that the resistors in Fig. 2.35a form a (delta) and the resistors in Fig. 2.35b form a Y (wye). If both of these configurations are connected at only three terminals a, b, and c, it would be very advantageous if an equivalence could be established between them. It is, in fact, possible to relate the resistances of one network to those of the other such that their terminal characteristics are the same. This relationship between the two network configurations is called the Y- transformation. The transformation that relates the resistances R1 , R2 , and R3 to the resistances Ra , Rb , and Rc is derived as follows. For the two networks to be equivalent at each corresponding pair of terminals, it is necessary that the resistance at the corresponding terminals be equal (e.g., the resistance at terminals a and b with c open-circuited must be the same for both networks). b (b) irwin02_025-100hr.qxd 62 30-06-2010 CHAPTER 2 13:14 Page 62 RESISTIVE CIRCUITS Therefore, if we equate the resistances for each corresponding set of terminals, we obtain the following equations: Rab = Ra + Rb = R2 AR1 + R3 B R2 + R1 + R3 R3 AR1 + R2 B Rbc = Rb + Rc = R3 + R1 + R2 Rca = Rc + Ra = 2.27 R1 AR2 + R3 B R1 + R2 + R3 Solving this set of equations for Ra , Rb , and Rc yields Ra = R1 R2 R1 + R2 + R3 Rb = R2 R3 R1 + R2 + R3 Rc = R1 R3 R1 + R2 + R3 2.28 Similarly, if we solve Eq. (2.27) for R1 , R2 , and R3 , we obtain R1 = Ra Rb + Rb Rc + Ra Rc Rb R2 = Ra Rb + Rb Rc + Ra Rc Rc R3 = Ra Rb + Rb Rc + Ra Rc Ra 2.29 Equations (2.28) and (2.29) are general relationships and apply to any set of resistances connected in a Y or . For the balanced case where Ra=Rb=Rc and R1=R2=R3 , the equations above reduce to 1 R 3 ¢ 2.30 R¢ = 3RY 2.31 RY = and It is important to note that it is not necessary to memorize the formulas in Eqs. (2.28) and (2.29). Close inspection of these equations and Fig. 2.35 illustrates a definite pattern to the relationships between the two configurations. For example, the resistance connected to point a in the wye (i.e., Ra) is equal to the product of the two resistors in the that are connected to point a divided by the sum of all the resistances in the delta. Rb and Rc are determined in a similar manner. Similarly, there are geometrical patterns associated with the equations for calculating the resistors in the delta as a function of those in the wye. Let us now examine the use of the delta ∆ wye transformation in the solution of a network problem. irwin02_025-100hr.qxd 30-06-2010 13:14 Page 63 SECTION 2.7 W Y E ∆ D E LTA T R A N S F O R M AT I O N S Given the network in Fig. 2.36a, let us find the source current IS . IS ± – EXAMPLE 2.26 IS 12 k Figure 2.36 18 k Circuits used in Example 2.26. 6 k 6 k ± – 12 V 12 V 3 k 2 k 4 k 9 k 4 k (a) 9 k (b) Note that none of the resistors in the circuit are in series or parallel. However, careful examination of the network indicates that the 12k-, 6k-, and 18k-ohm resistors, as well as the 4k-, 6k-, and 9k-ohm resistors each form a delta that can be converted to a wye. Furthermore, the 12k-, 6k-, and 4k-ohm resistors, as well as the 18k-, 6k-, and 9k-ohm resistors, each form a wye that can be converted to a delta. Any one of these conversions will lead to a solution. We will perform a delta-to-wye transformation on the 12k-, 6k-, and 18k-ohm resistors, which leads to the circuit in Fig. 2.36b. The 2k- and 4k-ohm resistors, like the 3k- and 9k-ohm resistors, are in series and their parallel combination yields a 4k-ohm resistor. Thus, the source current is SOLUTION IS = 12兾(6k + 4k) = 1.2 mA Learning Assessments E2.24 Determine the total resistance RT in the circuit in Fig. E2.24. ANSWER: RT = 34 k. 6 k 54 k 36 k 18 k RT 3 k 18 k 2 k Figure E2.24 E2.25 Find Vo in the network in Fig. E2.25. ANSWER: Vo = 24 V. 12 k 12 k 12 k 4 mA + 12 k Figure E2.25 63 12 k Vo – irwin02_025-100hr.qxd 64 30-06-2010 CHAPTER 2 13:14 Page 64 RESISTIVE CIRCUITS E2.26 Find I1 in Fig. E2.26. ANSWER: I1=-1.2 A. 18 18 18 3A 6 12 I1 Figure E2.26 2.8 Circuits with Dependent Sources In Chapter 1 we outlined the different kinds of dependent sources. These controlled sources are extremely important because they are used to model physical devices such as npn and pnp bipolar junction transistors (BJTs) and field-effect transistors (FETs) that are either metaloxide-semiconductor field-effect transistors (MOSFETs) or insulated-gate field-effect transistors (IGFETs). These basic structures are, in turn, used to make analog and digital devices. A typical analog device is an operational amplifier (op-amp). This device is presented in Chapter 4. Typical digital devices are random access memories (RAMs), read-only memories (ROMs), and microprocessors. We will now show how to solve simple one-loop and one-node circuits that contain these dependent sources. Although the following examples are fairly simple, they will serve to illustrate the basic concepts. Problem-Solving Strategy Circuits with Dependent Sources Step 1. When writing the KVL and/or KCL equations for the network, treat the dependent source as though it were an independent source. Step 2. Write the equation that specifies the relationship of the dependent source to the controlling parameter. Step 3. Solve the equations for the unknowns. Be sure that the number of linearly independent equations matches the number of unknowns. The following four examples will each illustrate one of the four types of dependent sources: current-controlled voltage source, current-controlled current source, voltagecontrolled voltage source, and voltage-controlled current source. EXAMPLE Let us determine the voltage Vo in the circuit in Fig. 2.37. 2.27 I1 3 k + Va=2000 I1 Figure 2.37 Circuit used in Example 2.27. 12 V ± – 5 k + Vo – irwin02_025-100hr.qxd 30-06-2010 13:14 Page 65 SECTION 2.8 CIRCUITS WITH DEPENDENT SOURCES Applying KVL, we obtain 65 SOLUTION -12 + 3kI1 - VA + 5kI1 = 0 where VA = 2000I1 and the units of the multiplier, 2000, are ohms. Solving these equations yields I1 = 2 mA Then Vo = (5 k)I1 = 10 V Given the circuit in Fig. 2.38 containing a current-controlled current source, let us find the voltage Vo . EXAMPLE 2.28 Figure 2.38 + 2 k 10 mA 3 k + 4 k Vo 4 Io Io – VS – SOLUTION Applying KCL at the top node, we obtain 10 * 10-3 + Circuit used in Example 2.28. VS VS + - 4Io = 0 2k + 4k 3k where Io = VS 3k Substituting this expression for the controlled source into the KCL equation yields 10-2 + VS VS 4VS + = 0 6k 3k 3k Solving this equation for VS , we obtain VS = 12 V The voltage Vo can now be obtained using a simple voltage divider; that is, Vo = c 4k d VS 2k + 4k = 8V The network in Fig. 2.39 contains a voltage-controlled voltage source. We wish to find Vo in this circuit. I EXAMPLE 2.29 3 k + 12 V ± – 2Vo 1 k + Vo Figure 2.39 – Circuit used in Example 2.29. irwin02_025-100hr.qxd 66 30-06-2010 CHAPTER 2 13:14 Page 66 RESISTIVE CIRCUITS SOLUTION Applying KVL to this network yields -12 + 3kI + 2Vo + 1kI = 0 where Vo = 1kI Hence, the KVL equation can be written as -12 + 3kI + 2kI + 1kI = 0 or I=2 mA Therefore, Vo = 1kI = 2V EXAMPLE 2.30 Figure 2.40 Example circuit containing a voltage-controlled current source. An equivalent circuit for a FET common-source amplifier or BJT common-emitter amplifier can be modeled by the circuit shown in Fig. 2.40a. We wish to determine an expression for the gain of the amplifier, which is the ratio of the output voltage to the input voltage. i1(t) R1 + + vi(t) ± – R2 vg(t) gm vg(t) R3 R4 – R5 vo(t) – (a) i1(t) R1 + + vi(t) ± – R2 vg(t) gm vg(t) RL vo(t) – – (b) SOLUTION Note that although this circuit, which contains a voltage-controlled current source, appears to be somewhat complicated, we are actually in a position now to solve it with techniques we have studied up to this point. The loop on the left, or input to the amplifier, is essentially detached from the output portion of the amplifier on the right. The voltage across R2 is vg(t), which controls the dependent current source. To simplify the analysis, let us replace the resistors R3 , R4 , and R5 with RL such that 1 1 1 1 = + + RL R3 R4 R5 Then the circuit reduces to that shown in Fig. 2.40b. Applying Kirchhoff’s voltage law to the input portion of the amplifier yields vi(t) = i1(t)AR1 + R2 B irwin02_025-100hr.qxd 30-06-2010 13:14 Page 67 SECTION 2.8 CIRCUITS WITH DEPENDENT SOURCES and vg(t)=i1(t)R2 Solving these equations for vg(t) yields R2 vi(t) R1 + R2 vg(t) = From the output circuit, note that the voltage vo(t) is given by the expression vo(t)=–gm vg(t)RL Combining this equation with the preceding one yields -gm RL R2 vi(t) R1 + R2 vo(t) = Therefore, the amplifier gain, which is the ratio of the output voltage to the input voltage, is given by vo(t) vi(t) = - gm RL R2 R1 + R2 Reasonable values for the circuit parameters in Fig. 2.40a are R1 = 100 , R2 = 1 k, gm = 0.04 S, R3 = 50 k, and R4 = R5 = 10 k. Hence, the gain of the amplifier under these conditions is vo(t) vi(t) = -(0.04)(4.545)A103 B(1)A103 B (1.1)A103 B = -165.29 Thus, the magnitude of the gain is 165.29. At this point it is perhaps helpful to point out again that when analyzing circuits with dependent sources, we first treat the dependent source as though it were an independent source when we write a Kirchhoff’s current or voltage law equation. Once the equation is written, we then write the controlling equation that specifies the relationship of the dependent source to the unknown variable. For instance, the first equation in Example 2.28 treats the dependent source like an independent source. The second equation in the example specifies the relationship of the dependent source to the voltage, which is the unknown in the first equation. Learning Assessments E2.27 Find Vo in the circuit in Fig. E2.27. I+VA – ANSWER: Vo = 12 V. –± 2VA + 4 k 6V Figure E2.27 ± – 8 k Vo – 67 irwin02_025-100hr.qxd 68 30-06-2010 CHAPTER 2 13:14 Page 68 RESISTIVE CIRCUITS ANSWER: Vo = 8 V. E2.28 Find Vo in the network in Fig. E2.28. 1 k⍀ + Vo — 2000 VS 6 k⍀ 2 mA + – 2 k⍀ Vo – Figure E2.28 E2.29 Find VA in Fig. E2.29. ANSWER: VA = -12 V. 2Vx 10 k⍀ 5 k⍀ +– 12 V VA + + – Vx – Figure E2.29 – + – 36 V + 5 k⍀ E2.30 Find V1 in Fig. E2.30. ANSWER: V1 = -32/3 V. – 4 k⍀ Vx + +– + 0.5 Vx + – – + V1 50 V 8 k⍀ – Figure E2.30 8 k⍀ 18 V E2.31 Find 1x in Fig. E2.31. ANSWER: Ix = -1.5 mA. Ix 2Ix 10 mA 2 k⍀ 3 mA 5 k⍀ 10 k⍀ Figure E2.31 E2.32 Find Vo in Fig. E2.32. ANSWER: Vo = 16 V. Ix 4 k⍀ Figure E2.32 + 6 mA 6 k⍀ 0.5Ix 12 k⍀ Vo – irwin02_025-100hr.qxd 30-06-2010 13:14 Page 69 SECTION 2.9 R E S I S TO R T E C H N O L O G I E S F O R E L E C T R O N I C M A N U FAC T U R I N G E2.33 If the power supplied by the 3-A current source in Fig. E2.33 is 12 W, find VS and the power supplied by the 10-V source. 3A 3⍀ Vs + – 69 ANSWER: VS = 42 V, -30 W. 4⍀ 4⍀ 6⍀ 5⍀ 10 V + – Figure E2.33 2.9 In addition to the resistors shown in Fig 2.1, three types are employed in the modern electronics industry: thick-film, thin-film, and silicon-diffused resistors. T H I C K- F I L M R E S I S TO R S Thick-film resistor components are found on all modern surface mount technology (SMT) printed circuit boards. They come in a variety of shapes, sizes, and values. A table of standard sizes for thick-film chip resistors is shown in Table 2.2, and some examples of surface mount thick-film ceramic resistors can be seen in Fig. 2.41. Thick-film resistors are considered “low-tech,” when compared with thin-film and silicon-diffused components, because they are manufactured using a screen printing process similar to that used with T-shirts. The screens utilized in thick-film manufacturing use a much finer mesh and are typically made of stainless steel for a longer lifetime. The paste used in screen printing resistors consists of a mixture of ruthenium oxides (RuO2) and glass. Resistor Technologies for Electronic Manufacturing TABLE 2.2 Thick-film chip resistor standard sizes SIZE CODE 0201 0402 0603 0805 1206 2010 2512 SIZE (MILS) 20 40 60 80 120 200 250 * * * * * * * 10 20 30 50 60 100 120 POWER RATING (WATTS) 1兾20 1兾16 1兾10 1兾8 1兾4 1兾2 1 Figure 2.41 A printed circuit board showing surface mount thick-film ceramic resistors. (Courtesy of Mike Palmer) Chip Resistors Chip Capacitors irwin02_025-100hr.qxd 70 30-06-2010 CHAPTER 2 13:14 Page 70 RESISTIVE CIRCUITS Figure 2.42 Thick-film chip resistor cross-section. Protective coating Electrode (Inner) Electrode (Between) Alumina substrate Thick film resistive element Electrode (Outer) Once the paste is screen printed, it is fired at temperatures around 850°C, causing the organic binders to vaporize and to allow the glass to melt and bind the metal and glass filler to the substrate. The substrates are typically 95% alumina ceramic. After firing, conductors are screen printed and fired to form the contacts used to solder the resistors. A second layer of glass is screen printed and fired to seal and protect the resistor. A cross-section of a typical thick-film resistor is shown in Fig. 2.42. Notice that the conductors are “wrapped” around the substrate to allow them to be soldered from the bottom or top and allow the solder to “wic” up the side to form a more reliable mechanical and electrical contact. Thick-film resistors have typical “as fired” tolerances of +兾- 10% to +兾- 20%. These wide tolerances are due to the fact that the screen printing process does not afford good geometry transfer or consistent thickness. To obtain a better tolerance (i.e., +兾- 0.5% to +兾- 1.0%), the resistors can be trimmed with a YAG laser to remove a portion of the resistor and change its value. The resistor is constantly measured during the cutting process to make sure the resistance is within the specified tolerance. T H I N - F I L M R E S I S TO R S Thin-film resistors are fabricated by depositing a thin layer (hundreds of angstroms, where one angstrom is one ten-billionth of a meter) of Tantalum Nitride (TaN) or Nichrome (NiCr) onto a silicon or highly polished alumina ceramic substrate. Using a photolithography process, the metal film is patterned and etched to form the resistor structure. Thin-film metals have a limited resistivity (the reciprocal of conductivity— a measure of a material’s ability to carry an electric current). This low resistivity limits the practical range of thin-film resistors due to the large areas required. Both TaN and NiCr have similar characteristics, but TaN is more chemically and thermally resistant and will hold up better to harsh environments. Sputtered metal thin films are continuous and virtually defect free, which makes them very stable, low-noise components that have negligible nonlinearity when compared with the more porous thick-film materials. Thin-film resistors are available in standard SMT packages, but are also available as wirebondable chips that can be directly patterned onto integrated circuits. A cross-sectional drawing of the thin-film chip on ceramic or silicon is shown in Fig 2.43. Because of the additional sophistication involved in fabrication, thin-film resistors are more expensive than thickfilm resistors. However, they have a number of important characteristics that make them the preferred devices for a number of microwave applications. Like thick-film resistors, these components can also be laser-trimmed to obtain a desired value within a specified tolerance. Since the sputtered metal film is extremely thin, the power requirement for the laser is very low, which in turn ensures that there will be minimal micro-cracking and therefore an increased level of stability. S I L I CO N - D I F F U S E D R E S I S TO R S Silicon-diffused resistors are part of virtually all integrated circuits (ICs). They are passive devices that are implemented to support or enhance the capabilities of active devices, such as transistors and diodes. Both passive and active devices are manufactured at the same time using the same technology (e.g., CMOS—complementary metal-oxide semiconductor). The resistors are made by diffusing a dopant, such as boron or phosphorus, into a silicon substrate at high temperature. This process is very irwin02_025-100hr.qxd 30-06-2010 13:14 SECTION 2.9 Page 71 R E S I S TO R T E C H N O L O G I E S F O R E L E C T R O N I C M A N U FAC T U R I N G Protective Coat Epoxy Figure 2.43 Thin-film chip resistor cross-section. Sputtered Resistive Metal Film Terminations (Nickel & Solder Plating) Electrode Thin Film High Purity Alumina Substrate expensive and is the reason silicon-diffused resistors cost more than thin- or thick-film resistors. A photo of an integrated silicon resistor is shown in Fig. 2.44. Notice that the resistor is completely integrated within a larger circuit, because it is not economically feasible to make discrete silicon-diffused resistors. Table 2.3 compares some of the characteristics of thickfilm, thin-film, and silicon-diffused resistors. Silicon resistors have a resistance range on the order of 5–6k ohms/sq. The term “ohms per square” means a dimensionless square area of resistive material, having an ohmic value equal Diffused Resistors Figure 2.44 Silicon-diffused resistors. 71 irwin02_025-100hr.qxd 72 30-06-2010 CHAPTER 2 13:14 Page 72 RESISTIVE CIRCUITS to the sheet resistivity of the material. For example, a 10-ohm sheet resistivity material would constitute a 10-ohm resistor whether the material was 1 mil by 1 mil or 1 inch by 1 inch. Dividing the length of the resistor by its width yields the number of squares, and multiplying the number of squares by the sheet resistance yields the resistance value. The total resistance values are limited because of the high cost of silicon area, and there are other circuit design techniques for implementing high-valued resistors through the judicious use of transistors. These devices suffer from large changes in value over temperature and some resistance change with applied voltage. As a result of these poor characteristics, thin-film resistors mounted on the surface of the silicon are used in place of diffused resistors in critical applications. TABLE 2.3 Characteristics of resistor types 2.10 Application Examples • APPLICATION EXAMPLE 2.31 CHARACTERISTIC THICK-FILM THIN-FILM SILICON-DIFFUSED Sheet resistance Sheet tolerance (as fired) Sheet tolerance (final) Relative cost 5 - 500k ohms兾sq /20% /1% Low 25 - 300 ohms兾sq /10% /1% High 5 - 6k ohms兾sq /2% N兾A Higher Throughout this book we endeavor to present a wide variety of examples that demonstrate the usefulness of the material under discussion in a practical environment. To enhance our presentation of the practical aspects of circuit analysis and design, we have dedicated sections, such as this one, in most chapters for the specific purpose of presenting additional applicationoriented examples. The eyes (heating elements) of an electric range are frequently made of resistive nichrome strips. Operation of the eye is quite simple. A current is passed through the heating element causing it to dissipate power in the form of heat. Also, a four-position selector switch, shown in Fig. 2.45, controls the power (heat) output. In this case the eye consists of two nichrome strips modeled by the resistors R1 and R2 , where R1 6 R2 . 1. How should positions A, B, C, and D be labeled with regard to high, medium, low, and off settings? 2. If we desire that high and medium correspond to 2000 W and 1200 W power dissipation, respectively, what are the values of R1 and R2 ? 3. What is the power dissipation at the low setting? SOLUTION Position A is the off setting since no current flows to the heater elements. In position B, current flows through R2 only, while in position C current flows through R1 only. Since R1 6 R2 , more power will be dissipated when the switch is at position C. Thus, position C is the medium setting, B is the low setting, and, by elimination, position D is the high setting. When the switch is at the medium setting, only R1 dissipates power, and we can write R1 as R1 = V2S 2302 = P1 1200 or R1 = 44.08 irwin02_025-100hr.qxd 30-06-2010 13:14 Page 73 SECTION 2.10 A P P L I C AT I O N E X A M P L E S 73 On the high setting, 2000 W of total power is delivered to R1 and R2 . Since R1 dissipates 1200 W, R2 must dissipate the remaining 800 W. Therefore, R2 is R2 = V2S 2302 = P2 800 or R2 = 66.13 Finally, at the low setting, only R2 is connected to the voltage source; thus, the power dissipation at this setting is 800 W. A Figure 2.45 B Simple resistive heater selector circuit. C D VS=230 V ± – R2 R1 Have you ever cranked your car with the headlights on? While the starter kicked the engine, you probably saw the headlights dim then return to normal brightness once the engine was running on its own. Can we create a model to predict this phenomenon? Yes, we can. Consider the conceptual circuit in Fig. 2.46a and the model circuit in Fig. 2.46b, which isolates just the battery, headlights, and starter. Note the resistor Rbatt . It is included to model several power loss mechanisms that can occur between the battery and the loads, that is, the headlights and starter. First, there are the chemical processes within the battery itself which are not 100% efficient. Second, there are the electrical connections at both the battery posts and the loads. Third, the wiring itself has some resistance, although this is usually so small that it is negligible. The sum of these losses is modeled by Rbatt , and we expect the value of Rbatt to be small. A reasonable value is 25 m. Next we address the starter. When energized, a typical automobile starter will draw between 90 and 120 A. We will use 100 A as a typical number. Finally, the headlights will draw much less current—perhaps only 1 A. Now we have values to use in our model circuit. +12 V- Vbatt 12 V (a) SOLUTION Figure 2.46 A conceptual (a) model and (b) circuit for examining the effect of starter current on headlight intensity. Rbatt Headlights Headlight switch APPLICATION EXAMPLE 2.32 ± – VL IHL Istart Ignition switch (b) • irwin02_025-100hr.qxd 74 30-06-2010 CHAPTER 2 13:14 Page 74 RESISTIVE CIRCUITS Assume first that the starter is off. By applying KCL at the node labeled VL , we find that the voltage applied to the headlights can be written as VL = Vbatt - IHL Rbatt Substituting our model values into this equation yields VL = 11.75 V—very close to 12 V. Now we energize the starter and apply KCL again: VL = Vbatt - AIHL + Istart BRbatt Now the voltage across the headlights is only 9.25 V. No wonder the headlights dim! How would corrosion or loose connections on the battery posts change the situation? In this case, we would expect the quality of the connection from battery to load to deteriorate, increasing Rbatt and compounding the headlight dimming issue. • APPLICATION EXAMPLE 2.33 A Wheatstone bridge circuit is an accurate device for measuring resistance. This circuit, shown in Fig. 2.47, is used to measure the unknown resistor Rx . The center leg of the circuit contains a galvanometer, which is a very sensitive device that can be used to measure current in the microamp range. When the unknown resistor is connected to the bridge, R3 is adjusted until the current in the galvanometer is zero, at which point the bridge is balanced. In this balanced condition R1 R2 = R3 Rx so that Figure 2.47 The Wheatstone bridge circuit. Rx = a R2 bR R1 3 R1 I1 R3 R2 I2 G IG Ix I3 Rx Engineers also use this bridge circuit to measure strain in solid material. For example, a system used to determine the weight of a truck is shown in Fig. 2.48a. The platform is supported by cylinders on which strain gauges are mounted. The strain gauges, which measure strain when the cylinder deflects under load, are connected to a Wheatstone bridge as shown in Fig. 2.48b. The strain gauge has a resistance of 120 under no-load conditions and changes value under load. The variable resistor in the bridge is a calibrated precision device. Weight is determined in the following manner. The ¢R3 required to balance the bridge represents the ¢ strain, which when multiplied by the modulus of elasticity yields the ¢ stress. The ¢ stress multiplied by the cross-sectional area of the cylinder produces the ¢ load, which is used to determine weight. Let us determine the value of R3 under no load when the bridge is balanced and its value when the resistance of the strain gauge changes to 120.24 under load. SOLUTION Using the balance equation for the bridge, the value of R3 at no load is R3 = a R1 bR R2 x irwin02_025-100hr.qxd 30-06-2010 13:14 Page 75 SECTION 2.11 DESIGN EXAMPLES 75 = a 100 b (120) 110 = 109.0909 Under load, the value of R3 is R3 = a 100 b (120.24) 110 = 109.3091 Therefore, the ¢R3 is ¢R3 = 109.3091 - 109.0909 = 0.2182 Platform Figure 2.48 Diagrams used in Example 2.33. Strain gauge (a) R1=100 ⍀ R2=110 ⍀ G R3 Strain gauge Rx (b) Most of this text is concerned with circuit analysis; that is, given a circuit in which all the components are specified, analysis involves finding such things as the voltage across some element or the current through another. Furthermore, the solution of an analysis problem is generally unique. In contrast, design involves determining the circuit configuration that will meet certain specifications. In addition, the solution is generally not unique in that there may be many ways to satisfy the circuit/performance specifications. It is also possible that there is no solution that will meet the design criteria. In addition to meeting certain technical specifications, designs normally must also meet other criteria, such as economic, environmental, and safety constraints. For example, if a circuit design that meets the technical specifications is either too expensive or unsafe, it is not viable regardless of its technical merit. At this point, the number of elements that we can employ in circuit design is limited primarily to the linear resistor and the active elements we have presented. However, as we progress through the text we will introduce a number of other elements (for example, the op-amp, capacitor, and inductor), which will significantly enhance our design capability. We begin our discussion of circuit design by considering a couple of simple examples that demonstrate the selection of specific components to meet certain circuit specifications. 2.11 Design Examples irwin02_025-100hr.qxd 76 • 30-06-2010 CHAPTER 2 13:14 Page 76 RESISTIVE CIRCUITS DESIGN EXAMPLE 2.34 SOLUTION An electronics hobbyist who has built his own stereo amplifier wants to add a back-lit display panel to his creation for that professional look. His panel design requires seven light bulbs—two operate at 12 V兾15 mA and five at 9 V兾5 mA. Luckily, his stereo design already has a quality 12-V dc supply; however, there is no 9-V supply. Rather than building a new dc power supply, let us use the inexpensive circuit shown in Fig. 2.49a to design a 12-V to 9-V converter with the restriction that the variation in V2 be no more than ;5%. In particular, we must determine the necessary values of R1 and R2 . First, lamps L1 and L2 have no effect on V2 . Second, when lamps L3–L7 are on, they each have an equivalent resistance of Req = V2 9 = = 1.8 k I 0.005 As long as V2 remains fairly constant, the lamp resistance will also be fairly constant. Thus, the requisite model circuit for our design is shown in Fig. 2.49b. The voltage V2 will be at its maximum value of 9 + 5% = 9.45 V when L3–L7 are all off. In this case R1 and R2 are in series, and V2 can be expressed by simple voltage division as V2 = 9.45 = 12 c R2 d R1 + R2 Figure 2.49 12-V to 9-V converter circuit for powering panel lighting. R1 12 V ± – + R2 V2 - L1 L2 L3 L4 L5 (a) R1 12 V ± – + R2 V2 - 1.8 k⍀ 1.8 k⍀ 1.8 k⍀ (b) 1.8 k⍀ 1.8 k⍀ L6 L7 irwin02_025-100hr.qxd 30-06-2010 13:14 Page 77 SECTION 2.11 DESIGN EXAMPLES 77 Rearranging the equation yields R1 = 0.27 R2 A second expression involving R1 and R2 can be developed by considering the case when L3–L7 are all on, which causes V2 to reach its minimum value of 9–5%, or 8.55 V. Now, the effective resistance of the lamps is five 1.8-k resistors in parallel, or 360 . The corresponding expression for V2 is R2兾兾360 V2 = 8.55 = 12 c R1 + AR2兾兾360B d which can be rewritten in the form 360R1 + 360 + R1 R2 12 = = 1.4 360 8.55 Substituting the value determined for R1兾R2 into the preceding equation yields R1 = 360[1.4 - 1 - 0.27] or R1 = 48.1 and so for R2 R2 = 178.3 Let’s design a circuit that produces a 5-V output from a 12-V input. We will arbitrarily fix the power consumed by the circuit at 240 mW. Finally, we will choose the best possible standard resistor values from Table 2.1 and calculate the percent error in the output voltage that results from that choice. The simple voltage divider, shown in Fig. 2.50, is ideally suited for this application. We know that Vo is given by Vo = Vin c DESIGN EXAMPLE 2.35 SOLUTION R2 d R1 + R2 which can be written as R1 = R2 c Vin - 1d Vo Since all of the circuit’s power is supplied by the 12-V source, the total power is given by P = V2in R1 + R2 0.24 R1 Using the second equation to eliminate R1 , we find that R2 has a lower limit of R2 (5)(12) Vo Vin = = 250 P 0.24 12 V Substituting these results into the second equation yields the lower limit of R1 , that is R1 = R2 c ± – + R2 Vo=5 V - Vin - 1 d 350 Vo Thus, we find that a significant portion of Table 2.1 is not applicable to this design. However, determining the best pair of resistor values is primarily a trial-and-error operation Figure 2.50 A simple voltage divider • irwin02_025-100hr.qxd 78 30-06-2010 CHAPTER 2 13:14 Page 78 RESISTIVE CIRCUITS that can be enhanced by using an Excel spreadsheet as shown in Table 2.4. Standard resistor values from Table 2.1 were entered into Column A of the spreadsheet for R2. Using the equation above, theoretical values for R1 were calculated using R1 = 1.4ⴢR2. A standard resistor value was selected from Table 2.1 for R1 based on the theoretical calculation in Column B. V0 was calculated using the simple voltage-divider equation, and the power absorbed by R1 and R2 was calculated in Column E. Note that a number of combinations of R1 and R2 satisfy the power constraint for this circuit. The power absorbed decreases as R1 and R2 increase. Let’s select R1 = 1800 and R2 = 1300 , because this combination yields an output voltage of 5.032 V that is closest to the desired value of 5 V. The resulting error in the output voltage can be determined from the expression 5.032 - 5 Percent error = c d 100% = 0.64% 5 It should be noted, however, that these resistor values are nominal, that is, typical values. To find the worst-case error, we must consider that each resistor as purchased may be as much as ;5% off the nominal value. In this application, since V0 is already greater than the target of 5 V, the worst-case scenario occurs when V0 increases even further, that is, R1 is 5% too low (1710 ) and R2 is 5% too high (1365 ). The resulting output voltage is 5.32 V, which yields a percent error of 6.4%. Of course, most resistor values are closer to the nominal value than to the guaranteed maximum/minimum values. However, if we intend to build this circuit with a guaranteed tight output error such as 5% we should use resistors with lower tolerances. How much lower should the tolerances be? Our first equation can be altered to yield the worst-case output voltage by adding a tolerance, ¢ , to R2 and subtracting the tolerance from R1. Let’s choose a worst-case output voltage of V0max = 5.25 V, that is, a 5% error: TABLE 2.4 Spreadsheet calculations for simple voltage divider A 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 R2 300 330 360 390 430 470 510 560 620 680 750 820 910 1000 1100 1200 1300 1500 1600 1800 2000 2200 2400 B C D E R1 theor 420 462 504 546 602 658 714 784 868 952 1050 1148 1274 1400 1540 1680 1820 2100 2240 2520 2800 3080 3360 R1 430 470 510 560 620 680 750 750 910 910 1000 1100 1300 1300 1500 1600 1800 2000 2200 2400 2700 3000 3300 Vo 4.932 4.950 4.966 4.926 4.914 4.904 4.857 5.130 4.863 5.132 5.143 5.125 4.941 5.217 5.077 5.143 5.032 5.143 5.053 5.143 5.106 5.077 5.053 Pabs 0.197 0.180 0.166 0.152 0.137 0.125 0.114 0.110 0.094 0.091 0.082 0.075 0.065 0.063 0.055 0.051 0.046 0.041 0.038 0.034 0.031 0.028 0.025 irwin02_025-100hr.qxd 30-06-2010 13:14 Page 79 SECTION 2.11 V0max = 5.25 = Vin c R2(1 + ¢) R1(1 - ¢) + R2(1 + ¢) d = 12 c 1300(1 + ¢) 1800(1 - ¢) + 1300(1 + ¢) DESIGN EXAMPLES 79 d The resulting value of ¢ is 0.037, or 3.7%. Standard resistors are available in tolerances of 10, 5, 2, and 1%. Tighter tolerances are available but very expensive. Thus, based on nominal values of 1300 and 1800 , we should utilize 2% resistors to ensure an output voltage error less than 5%. In factory instrumentation, process parameters such as pressure and flowrate are measured, converted to electrical signals, and sent some distance to an electronic controller. The controller then decides what actions should be taken. One of the main concerns in these systems is the physical distance between the sensor and the controller. An industry standard format for encoding the measurement value is called the 4–20 mA standard, where the parameter range is linearly distributed from 4 to 20 mA. For example, a 100 psi pressure sensor would output 4 mA if the pressure were 0 psi, 20 mA at 100 psi, and 12 mA at 50 psi. But most instrumentation is based on voltages between 0 and 5 V, not on currents. Therefore, let us design a current-to-voltage converter that will output 5 V when the current signal is 20 mA. The circuit in Fig. 2.51a is a very accurate model of our situation. The wiring from the sensor unit to the controller has some resistance, Rwire . If the sensor output were a voltage proportional to pressure, the voltage drop in the line would cause measurement error even if the sensor output were an ideal source of voltage. But, since the data are contained in the current value, Rwire does not affect the accuracy at the controller as long as the sensor acts as an ideal current source. As for the current-to-voltage converter, it is extremely simple—a resistor. For 5 V at 20 mA, we employ Ohm’s law to find R = DESIGN EXAMPLE 2.36 SOLUTION 5 = 250 0.02 The resulting converter is added to the system in Fig. 2.51b, where we tacitly assume that the controller does not load the remaining portion of the circuit. Note that the model indicates that the distance between the sensor and controller could be infinite. Intuitively, this situation would appear to be unreasonable, and it is. Losses that would take place over distance can be accounted for by using a more accurate model of the sensor, as shown in Fig. 2.52. The effect of this new sensor model can be seen from the Figure 2.51 equations that describe this new network. The model equations are The 4- to 20-mA control IS = Sensor model loop (a) block diagram, (b) with the current-tovoltage converter. VS VS + RS Rwire + 250 Sensor model Rwire I to V converter (a) Controller Rwire 250 ⍀ Controller (b) • irwin02_025-100hr.qxd 80 30-06-2010 CHAPTER 2 13:14 Page 80 RESISTIVE CIRCUITS Figure 2.52 VS A more accurate model for the 4- to 20-mA control loop. Improved sensor model Rwire IS Isignal RS Controller 250 ⍀ and Isignal = VS Rwire + 250 Combining these equations yields Isignal IS = 1 Rwire + 250 1 + RS Thus, we see that it is the size of RS relative to (Rwire + 250 ) that determines the accuracy of the signal at the controller. Therefore, we want RS as large as possible. Both the maximum sensor output voltage and output resistance, RS , are specified by the sensor manufacturer. We will revisit this current-to-voltage converter in Chapter 4. • DESIGN EXAMPLE 2.37 The network in Fig. 2.53 is an equivalent circuit for a transistor amplifier used in a stereo preamplifier. The input circuitry, consisting of a 2-mV source in series with a 500- resistor, models the output of a compact disk player. The dependent source, Rin , and Ro model the transistor, which amplifies the signal and then sends it to the power amplifier. The 10-k load resistor models the input to the power amplifier that actually drives the speakers. We must design a transistor amplifier as shown in Fig. 2.53 that will provide an overall gain of –200. In practice we do not actually vary the device parameters to achieve the desired gain; rather, we select a transistor from the manufacturer’s data books that will satisfy the required specification. The model parameters for three different transistors are listed as follows: Manufacturer’s transistor parameter values Part Number Rin (k⍀) Ro (k⍀) gm (mA/V) 1.0 2.0 8.0 50 75 80 50 30 20 1 2 3 Design the amplifier by choosing the transistor that produces the most accurate gain. What is the percent error of your choice? SOLUTION The output voltage can be written Vo = -gm VARo兾兾RL B irwin02_025-100hr.qxd 30-06-2010 13:14 Page 81 SUMMARY 81 Using voltage division at the input to find V, V = VS a Rin b Rin + RS Combining these two expressions, we can solve for the gain: AV = Vo Rin = -gm a b ARo兾兾RL B VS Rin + RS Using the parameter values for the three transistors, we find that the best alternative is transistor number 2, which has a gain error of Percent error = a 211.8 - 200 b * 100% = 5.9% 200 RS=500 ⍀ Figure 2.53 + VS=2 mV ± – Rin V gm V Ro RL=10 k⍀ - + Vo - Transistor amplifier circuit model. SUMMARY • ■ Ohm’s law V = IR ■ ■ The passive sign convention with Ohm’s law The current enters the resistor terminal with the positive voltage reference. The current-division rule The current is divided between two parallel resistors in reverse proportion to their resistance. ■ The equivalent resistance of a network of resistors Combine resistors in series by adding their resistances. Combine resistors in parallel by adding their conductances. The wye-to-delta and delta-to-wye transformations are also an aid in reducing the complexity of a network. ■ Short circuit Zero resistance, zero voltage; the current in the short is determined by the rest of the circuit. ■ Open circuit Zero conductance, zero current; the voltage across the open terminals is determined by the rest of the circuit. ■ Kirchhoff’s current law (KCL) The algebraic sum of the currents leaving (entering) a node is zero. ■ Kirchhoff’s voltage law (KVL) The algebraic sum of the voltages around any closed path is zero. ■ Solving a single-loop circuit Determine the loop current by applying KVL and Ohm’s law. ■ Solving a single-node-pair circuit Determine the voltage between the pair of nodes by applying KCL and Ohm’s law. ■ The voltage-division rule The voltage is divided between two series resistors in direct proportion to their resistance. irwin02_025-100hr.qxd 82 • 30-06-2010 CHAPTER 2 13:15 Page 82 RESISTIVE CIRCUITS PROBLEMS 2.1 Determine the current and power dissipated in the resistor 2.7 A model for a standard two D-cell flashlight is shown in Fig. P2.7. Find the power dissipated in the lamp. in Fig. P2.1. 1-⍀ lamp 9V ± – 12 ⍀ Figure P2.1 1.5 V 2.2 Determine the current and power dissipated in the resistors in Fig. P2.2. 1.5 V 2⍀ 12 V ± – 0.5 S Figure P2.7 Figure P2.2 2.8 An automobile uses two halogen headlights connected as 2.3 Determine the voltage across the resistor in Fig. P2.3 and the power dissipated. shown in Fig. P2.8. Determine the power supplied by the battery if each headlight draws 3 A of current. 12 ⍀ 2A Figure P2.3 2.4 Given the circuit in Fig. P2.4, find the voltage across each resistor and the power dissipated in each. + - 12 V 5⍀ Figure P2.8 6A 0.25 S Figure P2.4 2.9 Many years ago a string of Christmas tree lights was man- 2.5 In the network in Fig. P2.5, the power absorbed by Rx is 20 mW. Find R x. 2 mA ufactured in the form shown in Fig. P2.9a. Today the lights are manufactured as shown in Fig. P2.9b. Is there a good reason for this change? Rx Figure P2.5 (a) 2.6 In the network in Fig. P2.6, the power absorbed by Gx is 20 mW. Find Gx. 12 mA Figure P2.6 Gx (b) Figure P2.9 irwin02_025-100hr.qxd 30-06-2010 13:15 Page 83 PROBLEMS 2.10 Find I1 in the network in Fig. P2.10. 83 2.13 Find I1 in the circuit in Fig. P2.13. I1 6 mA 4 mA 20 mA I1 4 mA ± – Figure P2.10 12 mA 2 mA 2.11 Find I1 in the network in Fig. P2.11. Figure P2.13 I1 6 mA ± – 2 mA 2.14 Find Ix in the network in Fig. P2.14. Figure P2.11 2.12 Find I1 and I2 in the network in Fig. P2.12. I1 I2 8 mA 2Ix 3Ix 4 mA 12 mA Ix Figure P2.14 ± – 4 mA 2 mA Figure P2.12 2.15 Determine IL in the circuit in Fig. P2.15. 6 k⍀ 6 mA 3 Ix 3 mA Ix IL 2 k⍀ 3 k⍀ Figure P2.15 2.16 Find Io and I1 in the circuit in Fig. P2.16. 2.17 Find I1 in the network in Fig. P2.17. 5 mA 2Ix 4 mA Ix I1 I1 2 mA ±– 4 mA 2 mA ± – Figure P2.17 Io Figure P2.16 3 mA irwin02_025-100hr.qxd 84 30-06-2010 CHAPTER 2 13:15 Page 84 RESISTIVE CIRCUITS 2.18 Find Ix, Iy, and Iz in the network in Fig. P2.18. 2.22 In the network in Fig. P2.22, Find I1, I2 and I3 and show that KCL is satisfied at the boundary. Ix 3 mA 12 mA 4 mA I1 –± Iy I3 1 mA 4 mA Iz I2 2 mA 2 mA 4 mA Figure P2.18 Figure P2.22 2.19 Find I1 in the circuit in Fig. P2.19. 4Ix 2.23 Find Vbd in the circuit in Fig. P2.23. b a 4 mA + 4V - 6V + ± – 12 V 12 mA c ±– 2V - Ix I1 d Figure P2.19 Figure P2.23 2.20 Find I1 in the network in Fig. P2.20. 2.24 Find Vad in the network in Fig. P2.24. 2Ix 4 mA - 3V + 6 mA Ix 4V c b a - 2V + ± – ± – e + 3V - d 12 V 6 mA I1 Figure P2.24 Figure P2.20 2.21 Find I1, I2, and I3 in the network in Fig. P2.21. 2.25 Find Vfb and Vec in the circuit in Fig. P2.25. I3 a 12 mA Ix 2V + 4 mA I2 12 V c +1 V- d ±– 3V + 3V - 2Ix 2Ix I1 Figure P2.21 b + g -2 V+ Figure P2.25 f -1 V+ e irwin02_025-100hr.qxd 30-06-2010 13:15 Page 85 PROBLEMS 2.29 Find Vx and Vy in the circuit in Fig. P2.29. 2.26 Find Vae and Vcf in the circuit in Fig. P2.26. Vx – b a + c ±– - 4V + 9V + - 6V 5V - + f d 6V +– + 12 V + – + 6V Vy 4V – ± – 12 V 85 – ± – Figure P2.29 e 2.30 Find V1, V2 and V3 in the network in Fig. P2.30. Figure P2.26 + Vx 2.27 Given the circuit diagram in Fig. P2.27, find the following voltages: Vda, Vbh, Vgc, Vdi, Vfa, Vac, Vai, Vhf, Vfb, and Vdc. 8V a + - b - 12 V V2 2Vx + + + V1 6V V3 – – – 20 V f + – – +– + + c e + – – + d 6V 4V Figure P2.30 - 2.31 Find Vo in the network in Fig. P2.31. + + 16 V + 8V - - g + 12 V - h - V + x – 14 V - 4V + i 4⍀ 12 V + + 4 Vx + + 2⍀ ± – VA 2 VA – Figure P2.27 Vo – Figure P2.31 2.28 Find Vx and Vy in the circuit in Fig. P2.28. + + Vy 12 V – – 2.32 Find Vo in the circuit in Fig. P2.32. + Ix Vx 4⍀ 2 Ix 6V 6V Figure P2.28 – Vx + 12 ⍀ 12 V – +– + – + – + + - 4V Figure P2.32 + Vo – ± – Vx irwin02_025-100hr.qxd 86 30-06-2010 CHAPTER 2 13:15 Page 86 RESISTIVE CIRCUITS 2.33 The 10-V source absorbs 2.5 mW of power. Calculate 2.37 Find Vbd in the network in Fig. P2.37. Vba and the power absorbed by the dependent voltage source in Fig. P2.33. 3 k⍀ 2 k⍀ a 12 V 2 Vx 3 k⍀ 1 k⍀ ± – ± – 4V + - ± – 20 V c b a Vba d Vx 10 k⍀ Figure P2.37 + 5 k⍀ + b 10 V 2.38 Find Vx in the circuit in Fig. P2.38. 24 V Figure P2.33 ±– + 2.34 Find V1, V2, and V3 in the network in Fig. P2.34. Vx 4 k⍀ 6 k⍀ + – – +– 4V 12 V + – + – 4Vx V2 6V 6V + + + Vx V3 – – ± – + V1 ± – 8V Figure P2.38 2.39 Find Vab in the network in Fig. P2.39. Figure P2.34 6V 2⍀ ±– a 2.35 The 10-V source in Fig. P.2.35 is supplying 50 W. 3⍀ Determine R1. 4⍀ 9V 6⍀ 10 V + – R1 4R1 Figure P2.35 –± b Figure P2.39 2.40 Find Vx and the power supplied by the 15-V source in the circuit in Fig. P2.40. 2.36 Find V1 and V2 in Fig. P2.36. 5 k⍀ 2A + 10 ⍀ V2 Vx 6 k⍀ ±– 4 k⍀ 5⍀ Figure P2.36 15 V V1 - + 2 k⍀ –± 5⍀ - 10 V Figure P2.40 25 V 8 k⍀ ± – irwin02_025-100hr.qxd 30-06-2010 13:15 Page 87 PROBLEMS 87 2.45 The 100-V source in the circuit in Fig. P2.45 is 2.41 Find V1 in the network in Fig. P2.41. supplying 200 W. Solve for V2. 30 ⍀ - Vx + + 10 k⍀ ± – 5⍀ 5 k⍀ ± – V1 25 V 5⍀ Vx + – 100 V 4 + V 2 – 20 ⍀ Figure P2.41 40 ⍀ Figure P2.45 2.42 Find the power supplied by each source, including the dependent source, in Fig. P2.42. 3 mA 10 k⍀ 2.46 Find the value of V2 in Fig. P2.46 such that V1 = 0. 10 ⍀ 5 k⍀ 20 ⍀ + 15 V + – + – Vx 20 V 3 Vx + – + V 2 – V1 – 3 k⍀ Figure P2.46 1 k⍀ Figure P2.42 2.47 Find Io in the network in Fig. P2.47. 2.43 Find the power absorbed by the dependent voltage source in the circuit in Fig. P2.43. 2 Vx 10 k⍀ 12 mA 5 k⍀ 2 k⍀ ±– 20 V 3 k⍀ Io – ± ± – 6 k⍀ 10 V Figure P2.47 Vx 2.48 Find Io in the network in Fig. P2.48. 2 k⍀ 6 k⍀ 10 k⍀ 12 mA 12 k⍀ 3 k⍀ 12 k⍀ Io Figure P2.43 Figure P2.48 2.44 Find the power absorbed by the dependent source in the circuit in Fig. P2.44. 4 k⍀ 2.49 Find the power supplied by each source in the circuit in 10 k⍀ ±– Fig. P2.49. 60 V 20 V ± – ± – Ix 6 k⍀ Figure P2.44 2000 Ix 4 mA 1 k⍀ 10 k⍀ Figure P2.49 2 k⍀ 2 mA 5 k⍀ irwin02_025-100hr.qxd 88 30-06-2010 CHAPTER 2 13:15 Page 88 RESISTIVE CIRCUITS 2.52 Find Io in the network in Fig. P2.52. 2.50 Find the current IA in the circuit in Fig. P2.50. 7 mA 4 k⍀ 1 k⍀ 3 mA 2 k⍀ 8⍀ 5 k⍀ IA 3 Vx 5A 6 k⍀ + Figure P2.50 4⍀ 2.51 Find Io in the network in Fig. P2.51. Io Vx - Figure P2.52 4 k⍀ 12 k⍀ 2.53 Determine IL in the circuit in Fig. P2.53. Ix Io 6 k⍀ 12 mA + 3 k⍀ V1 6 mA 6 k⍀ 2 k⍀ 3 Ix - Figure P2.51 Figure P2.53 2.54 Find the power absorbed by the dependent source in the network in Fig. P2.54. + 2 VA 4 k⍀ 1 k⍀ 2 k⍀ VA - Figure P2.54 2.55 Find R AB in the circuit in Fig. P2.55. A 9 k⍀ RAB 2 k⍀ 2 k⍀ 12 k⍀ 4 k⍀ B Figure P2.55 2.56 Find RAB in the network in Fig. P2.56. A 2 k⍀ 1 k⍀ RAB B Figure P2.56 4 k⍀ 3 k⍀ 6 k⍀ 2 k⍀ 3 mA IL 3 mA 5 k⍀ 3 k⍀ irwin02_025-100hr.qxd 30-06-2010 13:15 Page 89 PROBLEMS 2.57 Find R AB in the circuit in Fig. P2.57. 2.61 Find R AB in the circuit in Fig. P2.61. A A 2 k⍀ 2 k⍀ 6⍀ 1 k⍀ 12 ⍀ RAB 2 k⍀ 2 k⍀ 89 2 k⍀ 2 k⍀ 6⍀ 6⍀ 9⍀ 9⍀ 1 k⍀ RAB 1 k⍀ 12 ⍀ B Figure P2.57 2.58 Find R AB in the network in Fig. P2.58. B 9⍀ A Figure P2.61 4⍀ 6⍀ 2.62 Find R AB in the network in Fig. P2.62. RAB 8⍀ A 10 ⍀ 6⍀ 6⍀ B 6⍀ 12 ⍀ 2⍀ RAB Figure P2.58 9⍀ 2.59 Find R AB in the circuit in Fig. P2.59. A 6⍀ 6⍀ 6⍀ 4⍀ 12 ⍀ B 14 ⍀ Figure P2.62 RAB 10 ⍀ 2⍀ 6⍀ 2.63 Find the equivalent resistance Req in the network in Fig. P2.63. B Figure P2.59 12 ⍀ 2.60 Find R AB in the network in Fig. P2.60. 12 ⍀ A 12 ⍀ RAB 6⍀ 3⍀ 12 ⍀ 12 ⍀ 2⍀ 3⍀ B Figure P2.60 Req 4⍀ 12 ⍀ 2⍀ Figure P2.63 12 ⍀ 12 ⍀ 6⍀ irwin02_025-100hr.qxd 90 30-06-2010 CHAPTER 2 13:15 Page 90 RESISTIVE CIRCUITS 2.64 Find the equivalent resistance looking in at terminals a-b in the circuit in Fig. P2.64. 8⍀ 12 ⍀ 10 ⍀ 4⍀ 10 ⍀ 8⍀ 8⍀ 4⍀ 12 ⍀ 18 ⍀ 8⍀ a b 8⍀ 8⍀ 10 ⍀ 5⍀ 6⍀ 9⍀ 4⍀ 6⍀ 20 ⍀ Figure P2.64 2.65 Given the resistor configuration shown in Fig. P2.65, 2.67 Find I1 and V0 in the circuit in Fig. P2.67. find the equivalent resistance between the following sets of terminals: (1) a and b, (2) b and c, (3) a and c, (4) d and e, (5) a and e, (6) c and d, (7) a and d, (8) c and e, (9) b and d, and (10) b and e. 2 k⍀ 12 V a + 8 k⍀ ± – 6 k⍀ 4 k⍀ I1 10 ⍀ 5⍀ Vo - 5⍀ Figure P2.67 d b e 4⍀ 12 ⍀ 4⍀ 2.68 Find I1 and V0 in the circuit in Fig. P2.68. c I1 Figure P2.65 6V 2.66 Seventeen possible equivalent resistance values may be obtained using three resistors. Determine the seventeen different values if you are given resistors with standard values: 47 , 33 , and 15 . ± – + 2 k⍀ 12 k⍀ 4 k⍀ Vo - Figure P2.68 irwin02_025-100hr.qxd 30-06-2010 13:15 Page 91 91 PROBLEMS 2.69 Find Vab and Vdc in the circuit in Fig. P2.69. a Vab + - 2⍀ 2.74 Calculate Vab in Fig. P2.74. 20 ⍀ 5⍀ 75 V 20 V 4⍀ ± – + + – b 30 ⍀ 15 ⍀ a 20 ⍀ 3⍀ d 8⍀ 18 ⍀ Vab 1⍀ Vdc 12 ⍀ b Figure P2.74 - c 2⍀ 2.75 Calculate VAB in Fig. P2.75. Figure P2.69 B 2.70 Find V 1 and IA in the network in Fig. P2.70. VAB 4⍀ 16 k⍀ 6⍀ + ± – 15 V 10 k⍀ 8 k⍀ IA V1 A - 4⍀ 2⍀ 4⍀ 6⍀ 2⍀ 4⍀ 2A 2⍀ Figure P2.75 Figure P2.70 2.76 Calculate Vab and V1 in Fig. P2.76. 2.71 Find Io in the network in Fig. P2.71. Vab a 6 k⍀ 12 mA 12 k⍀ 12 k⍀ 2⍀ 12 k⍀ Io b 6⍀ 15 ⍀ 6A 4⍀ 6⍀ Figure P2.71 V1 4⍀ 4⍀ 2.72 Determine Io in the circuit in Fig. P2.72. Figure P2.76 6 k⍀ 12 k⍀ 4 k⍀ 4 k⍀ 2 k⍀ 2.77 Calculate VAB in Fig. P2.77. 16 k⍀ - 12 V + VAB Io A B 8 k⍀ 6 k⍀ 6 k⍀ Figure P2.72 4 k⍀ 12 k⍀ 2 k⍀ 2.73 Determine Vo in the network in Fig. P2.73. 5 k⍀ 18 mA 3 k⍀ 1 k⍀ + 30 mA – + Vo - Figure P2.73 Figure P2.77 18 V irwin02_025-100hr.qxd 92 30-06-2010 CHAPTER 2 13:15 Page 92 RESISTIVE CIRCUITS 2.82 If Io = 5 mA in the circuit in Fig. P2.82, find IS. 2.78 Calculate VAB and I1 in Fig. P2.78. VAB A Io = 5 mA B 8 k⍀ 5 k⍀ 4 k⍀ IS 4 k⍀ 2 k⍀ 3 k⍀ 12 k⍀ 6 k⍀ 4 k⍀ 10 k⍀ + – 6 k⍀ 40 V Figure P2.82 2.83 If Io = 2 mA in the circuit in Fig. P2.83, find VS. I1 1 k⍀ Figure P2.78 VS ± – 2.79 Calculate VAB and I1 in Fig. P2.79. 3 k⍀ 12 k⍀ 6 k⍀ Io 7 k⍀ I1 Figure P2.83 6 k⍀ 2 k⍀ 2.84 Find the value of VS in the network in Fig. P2.84 such that the power supplied by the current source is 0. A 6 k⍀ 6 k⍀ 40 V 8 k⍀ + – 8 k⍀ 3⍀ 8⍀ VAB ± – 3 k⍀ 18 V B VS ± – 3A 2⍀ Figure P2.79 6⍀ Figure P2.84 2.80 Find Vab in Fig. P2.80. 2.85 In the network in Fig. P2.85, Vo = 6 V. Find IS. 20 k⍀ 30 k⍀ 3 k⍀ 100 V + – – Vab + IS 1 k⍀ + Vo 7 k⍀ 2 k⍀ 2 k⍀ 100 k⍀ 50 k⍀ Figure P2.85 Figure P2.80 2.86 Find the value of V1 in the network in Fig. P2.86 such that Va = 0. 2.81 If Vo = 4 V in the network in Fig. P2.81, find VS. 8V 8 k⍀ + + VS ± – 2⍀ 4 k⍀ - Figure P2.81 Va Vo = 4 V V1 ± – - Figure P2.86 2⍀ + 2⍀ 4⍀ 2⍀ V1 ± – irwin02_025-100hr.qxd 30-06-2010 13:15 Page 93 PROBLEMS 2.91 If V2 = 4 V in Fig. P2.91, calculate Vx . 2.87 If V1 = 5 V in the circuit in Fig. P2.87, find IS. 3⍀ V1 = 5 V + - 4⍀ + 5⍀ IS 4 k⍀ 93 6 k⍀ 3 k⍀ 24 V + – 2⍀ I1 – 16 ⍀ Figure P2.87 V2 2⍀ 15 ⍀ 2⍀ 2 I1 +– Vx 2.88 In the network in Fig. P2.88, V1 = 12 V. Find VS. Figure P2.91 4 k⍀ + 2 k⍀ VS ± – V1 - 6 k⍀ 1 k⍀ 4 k⍀ 3 k⍀ 2.92 Find the value of IA in the network in Fig. P2.92. Figure P2.88 12 V –+ 2.89 Given that Vo = 4 V in the network in Fig. P2.89, find VS. + VS 3 k⍀ 2⍀ 2⍀ 2⍀ ±– 3 k⍀ 6V 2⍀ + 2 k⍀ 2 mA 12 k⍀ IA + – Vo=4 V 1 k⍀ 4V Figure P2.92 - Figure P2.89 2.93 Find the value of IA in the circuit in Fig. P2.93. 2.90 If VR = 15 V, find VX in Fig. P2.90. 4⍀ 4⍀ 3⍀ 3A 2A 8V 4⍀ IA 4⍀ + V x – + + 5⍀ VR + – 4⍀ + Vo = 12 V 4V – – Figure P2.90 6⍀ 4⍀ Figure P2.93 irwin02_025-100hr.qxd 94 30-06-2010 CHAPTER 2 13:15 Page 94 RESISTIVE CIRCUITS 2.94 Find in value of the current source IA in the network in 2.98 Given Io = 2 mA in the circuit in Fig. P2.98, find IA. Fig. P2.94. + – 2⍀ 3⍀ 6V 4⍀ ± – 6V 1 k⍀ 2 k⍀ 6V 2⍀ ±– + 2A + – 2⍀ IA IA 2⍀ 6V 4V 1 k⍀ 2 k⍀ 1 k⍀ Io – Figure P2.94 Figure P2.98 2.95 Given Vo = 12 V, find the value of IA in the circuit in Fig. P2.95. Ix + – 2⍀ 9⍀ 2 Ix 4⍀ 2.99 Given Io = 2 mA in the network in Fig. P2.99, find VA. 7⍀ + 2A – + 4⍀ IA Vo 4⍀ 12 V ± VA – 6 mA – 1 k⍀ 6V + Figure P2.95 1 k⍀ 2.96 Find the value of Vx in the network in Fig. P2.96, such 1 k⍀ that the 5-A current source supplies 50 W. 2 k⍀ 2 k⍀ Io 2⍀ ±– Figure P2.99 Vx 4⍀ 4⍀ 5V ± – 2⍀ 2⍀ 5A 2.100 Given Vo in the network in Fig. P2.100, find IA. Figure P2.96 2.97 The 5-A current source in Fig. P2.97 supplies 150 W. 15 ⍀ IA 1 k⍀ Calculate VA. ±– 2⍀ 5⍀ ±– 25 V VA + - 4⍀ 1 k⍀ 1 k⍀ 5⍀ 5A 6V ± – + 12 V 2 k⍀ 1 k⍀ Vo=4 V - 2⍀ Figure P2.97 Figure P2.100 irwin02_025-100hr.qxd 30-06-2010 13:15 Page 95 95 PROBLEMS 2.101 Find the value of Vx in the circuit in Fig. P2.101 such 2.104 Given that V1 = 4 V, find VA and RB in the circuit in that the power supplied by the 5-A source is 60 W. Fig. P2.104. Vx VA ±– 16 V + - ±– 4 mA 1⍀ 1⍀ RB 4 k⍀ 3A 8V 6 k⍀ 5 mA –± + V1 1 k⍀ 4⍀ 2 k⍀ 2 k⍀ - Figure P2.104 5V ± – 2⍀ 2⍀ 5A 2.105 Find the power absorbed by the network in Fig. P2.105. 6 k⍀ Figure P2.101 6 k⍀ 21 V ±– 12 k⍀ 2 k⍀ 18 k⍀ 2.102 The 3-A current source in Fig. P2.102 is absorbing 12 W. Determine R. Figure P2.105 2.106 Find the value of g in the network in Fig. P2.106 1⍀ 2⍀ 12 V such that the power supplied by the 3-A source is 20 W. Ix + – 1⍀ 3A 2⍀ 3A glx R 2⍀ 2⍀ Figure P2.106 Figure P2.102 2.107 Find the power supplied by the 24-V source in the circuit in Fig. P2.107. 2.103 If the power supplied by the 50-V source in Fig. P2.103 12 k⍀ is 100 W, find R. 5⍀ 2⍀ 12 k⍀ ±– 12 k⍀ 12 k⍀ 24 V 50 V + – Figure P2.103 2A R 12 k⍀ Figure P2.107 12 k⍀ irwin02_025-100hr.qxd 96 30-06-2010 CHAPTER 2 13:15 Page 96 RESISTIVE CIRCUITS 2.112 Find Vo in the network in Fig. P2.112. 2.108 Find Io in the circuit in Fig. P2.108. I 12 ⍀ 24 V ± – 12 ⍀ 8⍀ + 2 k⍀ 12 ⍀ + 2 Vo ± – Vo 4 k⍀ 24 V - 14 ⍀ Figure P2.112 Io Figure P2.108 2.113 Find Io in the circuit in Fig. P2.113. Vo 2.109 Find Io in the circuit in Fig. P2.109. 2⍀ Io 2Ix 3⍀ 9⍀ 36 V 2⍀ ± – 1⍀ 6A Io 12 ⍀ Ix 4⍀ Figure P2.113 12 ⍀ 5⍀ 18 ⍀ 2.114 Find Io in the circuit in Fig. P2.114. Figure P2.109 Vo Io 2.110 Determine the value of Vo in the network in Fig. P2.110. 2⍀ 12 k⍀ + 4A 2 Vx 2A + 6 k⍀ 1⍀ 18 k⍀ 4 k⍀ – Vo 6 k⍀ 12 V Vx Figure P2.114 ± – - 2.115 Find Vo in the circuit in Fig. P2.115. Figure P2.110 Vo Iy 2.111 Find Vo in the circuit in Fig. P2.111. 2⍀ IS + 3 k⍀ 12 V ± – + 2000 IS 5 k⍀ 6A Vo 2⍀ + 1⍀ Vx – Figure P2.111 2Iy Vx 2 Figure P2.115 irwin02_025-100hr.qxd 30-06-2010 13:15 Page 97 PROBLEMS 2.118 Find I1, I2, and I3 in the circuit in Fig. P2.118. 2.116 Find Vx in the network in Fig. P2.116. I3 Vo 6⍀ Ix 1⍀ 3⍀ 4 I1 4A 2Ix + 3⍀ 6⍀ 97 2⍀ I2 ± – 8⍀ 2A Vx ± – 24 V I1 – . Figure P2.118 Figure P2.116 2.119 Find Io in the network in Fig. P2.119. 2.117 Find Vo in the network in Fig. P2.117. Vo – 1⍀ + 2Vy 8⍀ Vx 2⍀ 4Vx 6A + 2⍀ 6⍀ 5A + 2⍀ Vy 4⍀ Io – Vx - Figure P2.119 Figure P2.117 2.120 A typical transistor amplifier is shown in Fig. P2.120. Find the amplifier gain G (i.e., the ratio of the output voltage to the input voltage). 100 ⍀ 4 k⍀ + VS=250 mV ± – 5 k⍀ + 500 ⍀ Ib 300 ⍀ 4*105 Ib Vo - Figure P2.120 2.121 Find the value of k in the network in Fig. P2.121, such that the power supplied by the 6-A source is 108 W. 4⍀ 6A 6⍀ 12 ⍀ k Io 6⍀ 3⍀ Io Figure P2.121 3 Vx irwin02_025-100hr.qxd 98 30-06-2010 CHAPTER 2 13:15 Page 98 RESISTIVE CIRCUITS 2.124 The power supplied by the 2-A current source in 2.122 Find the power supplied by the dependent current source in Fig. P2.122. Fig. P2.124 is 50 W, calculate k. 5⍀ 10 ⍀ 50 V + 5⍀ 2A Vx + - I1 2⍀ 4⍀ 5⍀ 2A 2⍀ kI1 0.4Vx Figure P2.124 – 2.125 Given the circuit in Fig. P2.125, solve for the value 10 ⍀ of k. 30 k⍀ V2 + Figure P2.122 – 30 k⍀ 50 V +– 2.123 If the power absorbed by the 10-V source in Fig. P2.123 is 40 W, calculate IS. 6⍀ 4⍀ 10 k⍀ 5⍀ kV2 3 mA + 0.6Vx + – 10 ⍀ Is 15 ⍀ Vx 18 k⍀ + - 10 V 9 k⍀ - Figure P2.125 Figure P2.123 TYPICAL PROBLEMS FOUND ON THE FE EXAM 2FE-1 What is the power generated by the source in the 2FE-2 Find Vab in the circuit in Fig. 2PFE-2. network in Fig. 2PFE-1? a. -5 V a. 2.8 W b. 10 V b. 1.2 W c. 15 V c. 3.6 W d. -10 V d. 2.4 W a 5 k⍀ 10 ⍀ 6 k⍀ 120 V 18 k⍀ 4A 5⍀ 12 k⍀ ± – Vab 15 ⍀ 4 k⍀ 6 k⍀ 10 ⍀ Figure 2PFE-1 b Figure 2PFE-2 irwin02_025-100hr.qxd 30-06-2010 13:15 Page 99 99 TYPICAL PROBLEMS FOUND ON THE FE EXAM 2FE-3 If Req = 10.8 in the circuit in Fig. 2PFE-3, 2FE-6 Find the power supplied by the 40 V source in the cir- what is R2 ? cuit in Fig. 2PFE-6. a. 12 a. 120 W b. 20 b. 232 W c. 8 c. 212 W d. 18 d. 184 W 50 ⍀ 4⍀ Req R2 8⍀ 20 ⍀ 40 V ± – 25 ⍀ 100 ⍀ 3A ± – 100 V 2⍀ Figure 2PFE-6 Figure 2PFE-3 2FE-7 What is the current Io in the circuit in Fig. 2PFE-7? a. 0.84 mA 2FE-4 Find the equivalent resistance of the circuit in Fig. 2PFE-4 at the terminals A-B. b. -1.25 mA a. 4 k c. 2.75 mA b. 12 k d. -0.22 mA c. 8 k 3 k⍀ d. 20 k 4 k⍀ + A 12 k⍀ 12 k⍀ 6 k⍀ 12 k⍀ 6 k⍀ RAB 12 V 12 k⍀ 6 k⍀ 6 k⍀ 12 k⍀ 3 k⍀ 4 k⍀ 6 k⍀ Io B Figure 2PFE-7 Figure 2PFE-4 2FE-8 Find the voltage Vo in the network in Fig. 2PFE-8. a. 24 V 2FE-5 The 100 V source is absorbing 50 W of power in the network in Fig. 2PFE-5. What is R? b. 10 V c. 36 V a. 17.27 d. 12 V b. 9.42 1 k⍀ c. 19.25 2 k⍀ d. 15.12 3 k⍀ 10 ⍀ 24 mA 6 k⍀ 10 ⍀ 6 k⍀ R ± – 5A ± – 100 V 200 V 12 k⍀ Figure 2PFE-5 Figure 2PFE-8 + Vo - irwin02_025-100hr.qxd 100 30-06-2010 CHAPTER 2 13:15 Page 100 RESISTIVE CIRCUITS 2FE-9 What is the voltage Vo in the circuit in Fig. 2PFE-9? 2FE-10 Find the current Ix in Fig. 2PFE-10. a. 2 V a. 1兾2 A b. 8 V b. 5兾3 A c. 5 V c. 3兾2 A d. 12 V d. 8兾3 A 1⍀ 1⍀ 2A 2⍀ 4A 1⍀ + 3⍀ Vo Ix 12 V ± – 3⍀ 10 ⍀ 2⍀ Figure 2PFE-9 Figure 2PFE-10 8⍀ irwin03_101-155hr2.qxd 28-07-2010 11:31 Page 101 C H A PT E R NODAL AND LOOP ANALYSIS TECHNIQUES 3 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Be able to calculate all currents and voltages in circuits that contain multiple nodes and loops ■ Learn to employ Kirchhoff’s current law (KCL) to perform a nodal analysis to determine all the node voltages in a circuit ■ Learn to employ Kirchhoff’s voltage law (KVL) to perform a loop analysis to determine all the loop currents in a network ■ Be able to ascertain which of the two analysis techniques should be utilized to solve a particular problem Courtesy of UPI/Ed Turner/Boeing/NewsCom B Boeing Dreamliner Truly a dream come true, the Boeing 787 An active gust alleviation system that automatically adjusts Dreamliner brings big-jet ranges to mid-size airplanes. A pleasure wing flaps using sensor data of turbulence at the aircraft’s for both cross-country and intercontinental commercial travelers, nose improves flight control. this super-efficient airplane can carry 210 to 330 passengers on The design of the Dreamliner is based on fundamental laws long-range flights from nearly 3,000 miles to over 9,000 miles. that lead to algorithms most efficiently implemented on comput- Featuring composite materials, it boasts unmatched efficiency, ers. The two basic circuit analysis techniques described in this requiring 20% less fuel than its competitors. In various design chapter follow this same pattern. Nodal analysis is based on bal- stages for nearly six years, the final assembly plant opened in ancing currents coming into and out of nodes in the circuits. Mesh 2007 in Washington State and first flight occurred in late 2009. or loop analysis is based on balancing voltage increases and Inside its efficient hull lies state of the art electronics. drops around closed paths in the circuits. Both methods clearly Based on an open architecture, the Dreamliner has health- follow the fundamental laws introduced in Chapter 2. Unlike a monitoring systems that allow the airplane to self-monitor and branch-by-branch analysis that yields large numbers of simple report maintenance requirements to ground computer systems. equations, these two methods use network topology to provide a A wireless broadband link sends this real-time diagnostic data minimum number of equations. Just as the Dreamliner outstrips to technicians on the ground. The goal is safer operations its predecessors in efficiency and range, these all-encompassing with predicted mechanical problems and shorter repair times. techniques can easily handle more complex linear circuits. irwin03_101-155hr.qxd 102 30-06-2010 CHAPTER 3 13:12 Page 102 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.1 Nodal Analysis In a nodal analysis, the variables in the circuit are selected to be the node voltages. The node voltages are defined with respect to a common point in the circuit. One node is selected as the reference node, and all other node voltages are defined with respect to that node. Quite often this node is the one to which the largest number of branches are connected. It is commonly called ground because it is said to be at ground-zero potential, and it sometimes represents the chassis or ground line in a practical circuit. We will select our variables as being positive with respect to the reference node. If one or more of the node voltages are actually negative with respect to the reference node, the analysis will indicate it. In order to understand the value of knowing all the node voltages in a network, we consider once again the network in Fig. 2.32, which is redrawn in Fig. 3.1. The voltages, VS , Va , Vb , and Vc , are all measured with respect to the bottom node, which is selected as the reference and labeled with the ground symbol . Therefore, the voltage at node 1 is VS = 12 V with respect to the reference node 5; the voltage at node 2 is Va = 3 V with respect to the reference node 5, and so on. Now note carefully that once these node voltages are known, we can immediately calculate any branch current or the power supplied or absorbed by any element, since we know the voltage across every element in the network. For example, the voltage V1 across the leftmost 9-k resistor is the difference in potential between the two ends of the resistor; that is, V1 = VS - Va = 12 - 3 = 9V This equation is really nothing more than an application of KVL around the leftmost loop; that is, -VS + V1 + Va = 0 In a similar manner, we find that V3 = Va - Vb and V5 = Vb - Vc Then the currents in the resistors are VS - Va V1 = 9k 9k V3 Va - Vb I3 = = 3k 3k V5 Vb - Vc I5 = = 9k 9k I1 = In addition, Va - 0 6k Vb - 0 I4 = 4k I2 = since the reference node 5 is at zero potential. Va=3 V Figure 3.1 VS Circuit with known node voltages. 1 12 V V + 1- I1 2 9 k⍀ + ± – – 3 V Vb=— 2 3 Vc=— V 8 V V + 3- 3 + 5I3 3 k⍀ + I5 9 k⍀ + 4 6 k⍀ 3 k⍀ I2 5 4 k⍀ – I4 – irwin03_101-155hr.qxd 30-06-2010 13:12 Page 103 SECTION 3.1 N O D A L A N A LY S I S 103 Figure 3.2 Node m + vm R Circuit used to illustrate Ohm’s law in a multiple-node network. Node N + vN i - - Thus, as a general rule, if we know the node voltages in a circuit, we can calculate the current through any resistive element using Ohm’s law; that is, i = vm - vN R 3.1 as illustrated in Fig. 3.2. Now that we have demonstrated the value of knowing all the node voltages in a network, let us determine the manner in which to calculate them. In a nodal analysis, we employ KCL equations in such a way that the variables contained in these equations are the unknown node voltages of the network. As we have indicated, one of the nodes in an N-node circuit is selected as the reference node, and the voltages at all the remaining N - 1 nonreference nodes are measured with respect to this reference node. Using network topology, it can be shown that exactly N - 1 linearly independent KCL equations are required to determine the N - 1 unknown node voltages. Therefore, theoretically once one of the nodes in an N-node circuit has been selected as the reference node, our task is reduced to identifying the remaining N - 1 nonreference nodes and writing one KCL equation at each of them. In a multiple-node circuit, this process results in a set of N - 1 linearly independent simultaneous equations in which the variables are the N - 1 unknown node voltages. To help solidify this idea, consider once again Example 2.5. Note that in this circuit only four (i.e., any four) of the five KCL equations, one of which is written for each node in this fivenode network, are linearly independent. Furthermore, many of the branch currents in this example (those not contained in a source) can be written in terms of the node voltages as illustrated in Fig. 3.2 and expressed in Eq. (3.1). It is in this manner, as we will illustrate in the sections that follow, that the KCL equations contain the unknown node voltages. It is instructive to treat nodal analysis by examining several different types of circuits and illustrating the salient features of each. We begin with the simplest case. However, as a prelude to our discussion of the details of nodal analysis, experience indicates that it is worthwhile to digress for a moment to ensure that the concept of node voltage is clearly understood. At the outset it is important to specify a reference. For example, to state that the voltage at node A is 12 V means nothing unless we provide the reference point; that is, the voltage at node A is 12 V with respect to what? The circuit in Fig. 3.3 illustrates a portion of a network containing three nodes, one of which is the reference node. V1=4 V 1 V2=–2 V R2 R1 3 2 R3 Figure 3.3 An illustration of node voltages. irwin03_101-155hr.qxd 104 30-06-2010 CHAPTER 3 13:12 Page 104 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S The voltage V1 = 4 V is the voltage at node 1 with respect to the reference node 3. Similarly, the voltage V2 = -2 V is the voltage at node 2 with respect to node 3. In addition, however, the voltage at node 1 with respect to node 2 is ±6 V, and the voltage at node 2 with respect to node 1 is -6 V. Furthermore, since the current will flow from the node of higher potential to the node of lower potential, the current in R1 is from top to bottom, the current in R2 is from left to right, and the current in R3 is from bottom to top. These concepts have important ramifications in our daily lives. If a man were hanging in midair with one hand on one line and one hand on another and the dc line voltage of each line was exactly the same, the voltage across his heart would be zero and he would be safe. If, however, he let go of one line and let his feet touch the ground, the dc line voltage would then exist from his hand to his foot with his heart in the middle. He would probably be dead the instant his foot hit the ground. In the town where we live, a young man tried to retrieve his parakeet that had escaped its cage and was outside sitting on a power line. He stood on a metal ladder and with a metal pole reached for the parakeet; when the metal pole touched the power line, the man was killed instantly. Electric power is vital to our standard of living, but it is also very dangerous. The material in this book does not qualify you to handle it safely. Therefore, always be extremely careful around electric circuits. Now as we begin our discussion of nodal analysis, our approach will be to begin with simple cases and proceed in a systematic manner to those that are more challenging. Numerous examples will be the vehicle used to demonstrate each facet of this approach. Finally, at the end of this section, we will outline a strategy for attacking any circuit using nodal analysis. C I R C U I T S CO N TA I N I N G O N LY I N D E P E N D E N T C U R R E N T S O U R C E S Consider the network shown in Fig. 3.4. Note that this network contains three nodes, and thus we know that exactly N - 1 = 3 - 1 = 2 linearly independent KCL equations will be required to determine the N - 1 = 2 unknown node voltages. First, we select the bottom node as the reference node, and then the voltage at the two remaining nodes labeled v1 and v2 will be measured with respect to this node. The branch currents are assumed to flow in the directions indicated in the figures. If one or more of the branch currents are actually flowing in a direction opposite to that assumed, the analysis will simply produce a branch current that is negative. Applying KCL at node 1 yields -iA + i1 + i2 = 0 Using Ohm’s law (i=Gv) and noting that the reference node is at zero potential, we obtain [hint] Employing the passive sign convention. -iA + G1 Av1 - 0B + G2 Av1 - v2 B = 0 or KCL at node 2 yields AG1 + G2 Bv1 - G2 v2 = iA -i2 + iB + i3 = 0 or -G2 Av1 - v2 B + iB + G3 Av2 - 0B = 0 which can be expressed as -G2 v1 + AG2 + G3 Bv2 = -iB v1 Figure 3.4 A three-node circuit. 1 iA v2 i2 R2 2 R1 R3 iB i1 3 i3 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 105 SECTION 3.1 N O D A L A N A LY S I S 105 Therefore, the two equations for the two unknown node voltages v1 and v2 are AG1 + G2 Bv1 - G2 v2 = iA -G2 v1 + AG2 + G3 Bv2 = -iB 3.2 Note that the analysis has produced two simultaneous equations in the unknowns v1 and v2 . They can be solved using any convenient technique, and modern calculators and personal computers are very efficient tools for this application. In what follows, we will demonstrate three techniques for solving linearly independent simultaneous equations: Gaussian elimination, matrix analysis, and the MATLAB mathematical software package. A brief refresher that illustrates the use of both Gaussian elimination and matrix analysis in the solution of these equations is provided in the Problem-Solving Companion for this text. Use of the MATLAB software is straightforward, and we will demonstrate its use as we encounter the application. The KCL equations at nodes 1 and 2 produced two linearly independent simultaneous equations: -iA + i1 + i2 = 0 -i2 + iB + i3 = 0 The KCL equation for the third node (reference) is +iA - i1 - iB - i3 = 0 Note that if we add the first two equations, we obtain the third. Furthermore, any two of the equations can be used to derive the remaining equation. Therefore, in this N = 3 node circuit, only N - 1 = 2 of the equations are linearly independent and required to determine the N - 1 = 2 unknown node voltages. Note that a nodal analysis employs KCL in conjunction with Ohm’s law. Once the direction of the branch currents has been assumed, then Ohm’s law, as illustrated by Fig. 3.2 and expressed by Eq. (3.1), is used to express the branch currents in terms of the unknown node voltages. We can assume the currents to be in any direction. However, once we assume a particular direction, we must be very careful to write the currents correctly in terms of the node voltages using Ohm’s law. Suppose that the network in Fig. 3.4 has the following parameters: IA = 1 mA, R1 = 12 k, R2 = 6 k, IB = 4 mA, and R3 = 6 k. Let us determine all node voltages and branch currents. For purposes of illustration we will solve this problem using Gaussian elimination, matrix analysis, and MATLAB. Using the parameter values Eq. (3.2) becomes V1 c 1 1 1 + d - V2 c d = 1 * 10-3 12k 6k 6k -V1 c 1 1 1 d + V2 c + d = -4 * 10-3 6k 6k 6k where we employ capital letters because the voltages are constant. The equations can be written as V2 V1 = 1 * 10-3 4k 6k - V2 V1 + = -4 * 10-3 6k 3k Using Gaussian elimination, we solve the first equation for V1 in terms of V2 : 2 V1 = V2 a b + 4 3 EXAMPLE 3.1 SOLUTION irwin03_101-155hr.qxd 106 30-06-2010 CHAPTER 3 13:12 Page 106 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S This value is then substituted into the second equation to yield V2 -1 2 = -4 * 10-3 a V2 + 4 b + 6k 3 3k or V2 = -15 V This value for V2 is now substituted back into the equation for V1 in terms of V2 , which yields 2 V2 + 4 3 V1 = = -6 V The circuit equations can also be solved using matrix analysis. The general form of the matrix equation is GV = I where in this case 1 4k G= D 1 6k 1 6k V 1 * 10-3 T , V = B 1 R , and I = B R 1 V2 -4 * 10-3 3k - The solution to the matrix equation is V = G - 1I and therefore, 1 V1 4k B R = D V2 -1 6k -1 -1 6k 1 * 10-3 T B R 1 -4 * 10-3 3k To calculate the inverse of G, we need the adjoint and the determinant. The adjoint is 1 3k Adj G = D 1 6k 1 6k T 1 4k and the determinant is ∑G∑ = a = 1 1 -1 -1 ba b - a ba b 3k 4k 6k 6k 1 18k 2 Therefore, 1 V1 3k B R = 18k2 D V2 1 6k 1 6k 1 * 10-3 TB R 1 -4 * 10-3 4k 1 4 - 2 2 3k 6k = 18k2 D T 1 1 6k2 k2 = B -6 R -15 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 107 SECTION 3.1 N O D A L A N A LY S I S 107 Knowing the node voltages, we can determine all the currents using Ohm’s law: I1 = V1 -6 1 = = - mA R1 12k 2 I2 = -6 - (-15) V1 - V2 3 = = mA 6k 6k 2 I3 = V2 -15 5 = = - mA 6k 6k 2 and Fig. 3.5 illustrates the results of all the calculations. Note that KCL is satisfied at every node. V1=–6 V V2=–15 V 3 mA — 2 1 mA Figure 3.5 6 k⍀ 5 mA — 2 12 k⍀ 4 mA Circuit used in Example 3.1 6 k⍀ 1 mA — 2 Let us now examine the circuit in Fig. 3.6. The current directions are assumed as shown in the figure. We note that this network has four nodes. The node at the bottom of the circuit is selected as the reference node and labeled with the ground symbol. Since N = 4, N - 1 = 3 linearly independent KCL equations will be required to determine the three unknown nonreference node voltages labeled v1 , v2 , and v3 . At node 1, KCL yields i1 - iA + i2 - i3 = 0 or v3 - v1 v1 v1 - v2 - iA + = 0 R1 R2 R3 v1 a 1 1 1 1 1 + + b - v2 - v3 = iA R1 R2 R3 R2 R3 At node 2, KCL yields -i2 + i4 - i5 = 0 or - -v1 v3 - v2 v1 - v2 v2 + = 0 R2 R4 R5 1 1 1 1 1 + v2 a + + b - v3 = 0 R2 R2 R4 R5 R5 R3 Figure 3.6 i3 A four-node circuit. v1 i 2 1 R1 i1 v2 R2 R5 i 5 2 iA v3 3 R4 i4 iB irwin03_101-155hr.qxd 108 30-06-2010 CHAPTER 3 13:12 Page 108 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S At node 3, the equation is i3 + i5 + iB = 0 or v3 - v1 v3 - v2 + + iB = 0 R3 R5 -v1 1 1 1 1 - v2 + v3 a + b = -iB R3 R5 R3 R5 Grouping the node equations together, we obtain v1 a -v1 1 1 1 1 1 + + b - v2 - v3 = iA R1 R2 R3 R2 R3 1 1 1 1 1 + v2 a + + b - v3 = 0 R2 R2 R4 R5 R5 -v1 3.3 1 1 1 1 - v2 + v3 a + b = -iB R3 R5 R3 R5 Note that our analysis has produced three simultaneous equations in the three unknown node voltages v1 , v2 , and v3 . The equations can also be written in matrix form as 1 1 1 + + R1 R2 R3 1 F R2 1 R3 1 R2 1 1 1 + + R2 R4 R5 1 R5 - 1 R3 1 R5 - 1 1 + R3 R5 v1 iA V C v2 S = C 0 S v3 -iB 3.4 At this point it is important that we note the symmetrical form of the equations that describe the two previous networks. Eqs. (3.2) and (3.3) exhibit the same type of symmetrical form. The G matrix for each network is a symmetrical matrix. This symmetry is not accidental. The node equations for networks containing only resistors and independent current sources can always be written in this symmetrical form. We can take advantage of this fact and learn to write the equations by inspection. Note in the first equation of (3.2) that the coefficient of v1 is the sum of all the conductances connected to node 1 and the coefficient of v2 is the negative of the conductances connected between node 1 and node 2. The right-hand side of the equation is the sum of the currents entering node 1 through current sources. This equation is KCL at node 1. In the second equation in (3.2), the coefficient of v2 is the sum of all the conductances connected to node 2, the coefficient of v1 is the negative of the conductance connected between node 2 and node 1, and the right-hand side of the equation is the sum of the currents entering node 2 through current sources. This equation is KCL at node 2. Similarly, in the first equation in (3.3) the coefficient of v1 is the sum of the conductances connected to node 1, the coefficient of v2 is the negative of the conductance connected between node 1 and node 2, the coefficient of v3 is the negative of the conductance connected between node 1 and node 3, and the right-hand side of the equation is the sum of the currents entering node 1 through current sources. The other two equations in (3.3) are obtained in a similar manner. In general, if KCL is applied to node j with node voltage vj , the coefficient of vj is the sum of all the conductances connected to node j and the coefficients of the other node voltages Ae.g., vj - 1 , vj + 1 B are the negative of the sum of the conductances connected directly between these nodes and node j. The right-hand side of the equation is equal to the sum of the currents entering the node via current sources. Therefore, the left-hand side of the equation represents the sum of the currents leaving node j and the right-hand side of the equation represents the currents entering node j. irwin03_101-155hr.qxd 30-06-2010 13:12 Page 109 SECTION 3.1 Let us apply what we have just learned to write the equations for the network in Fig. 3.7 by inspection. Then given the following parameters, we will determine the node voltages using MATLAB: R1 = R2 = 2 k, R3 = R4 = 4 k, R5 = 1 k, iA = 4 mA, and iB = 2 mA. R1 N O D A L A N A LY S I S 109 EXAMPLE 3.2 Figure 3.7 Circuit used in Example 3.2. v2 iA v1 v3 R4 R2 R3 R5 iB The equations are SOLUTION 1 1 1 v1 a + b - v2(0) - v3 a b = -iA R1 R2 R1 -v1(0) + v2 a -v1 a 1 1 1 + b - v3 a b = iA - iB R3 R4 R4 1 1 1 1 1 b - v2 a b + v3 a + + b = 0 R1 R4 R1 R4 R5 which can also be written directly in matrix form as 1 1 + R1 R2 F 0 - 1 R1 0 1 1 + R3 R4 1 R4 1 R1 v1 -iA 1 V C v2 S = C iA - iB S R4 v3 0 1 1 1 + + R1 R4 R5 - Both the equations and the G matrix exhibit the symmetry that will always be present in circuits that contain only resistors and current sources. If the component values are now used, the matrix equation becomes 1 1 + 2k 2k F 0 - 1 2k 1 2k v1 -0.004 1 1 1 + V Cv2S = C 0.002S 4k 4k 4k v3 0 1 1 1 1 + + 4k 2k 4k 1k 0 - or C 0.001 0 -0.0005 v1 -0.004 0 0.0005 -0.00025 S C v2 S = C 0.002 S -0.0005 -0.00025 0.00175 v3 0 This equation is in the form of Gv = i. Therefore, v = G-1 i. Performing this operation yields the following voltages: v1 = -4.3636 V v2 = 3.6364 V v3 = -0.7273 V irwin03_101-155hr.qxd 110 30-06-2010 CHAPTER 3 13:12 Page 110 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S Learning Assessments E3.1 Write the node equations for the circuit in Fig. E3.1. V1 ANSWER: 1 1 V V = 4 * 10-3 , 4k 1 12k 2 V2 12 k⍀ 6 k⍀ 4 mA 6 k⍀ 2 mA -1 1 V1 + V2 = -2 * 10-3. 12k 4k Figure E3.1 ANSWER: V1 = 5.4286 V, V2 = 2.000 V, E3.2 Find all the node voltages in the network in Fig. E3.2 using MATLAB. 1 k⍀ V1 V2 2 k⍀ 4 mA V3 = 3.1429 V. 4 k⍀ V3 1 k⍀ 2 mA Figure E3.2 E3.3 Use nodal analysis to find Vo in Fig. E3.3. ANSWER: Vo = 2.79 V. 2 mA 6 k⍀ 2 k⍀ + 8 mA 3 k⍀ 6 k⍀ 1 k⍀ Vo – Figure E3.3 C I R C U I T S CO N TA I N I N G D E P E N D E N T C U R R E N T S O U R C E S The presence of a dependent source may destroy the symmetrical form of the nodal equations that define the circuit. Consider the circuit shown in Fig. 3.8, which contains a current-controlled current source. The KCL equations for the nonreference nodes are io + v1 v1 - v2 + = 0 R1 R2 and v2 - v1 + io - iA = 0 R2 where io = v2兾R3 . Simplifying the equations, we obtain irwin03_101-155hr.qxd 30-06-2010 13:12 Page 111 SECTION 3.1 N O D A L A N A LY S I S 111 AG1 + G2 Bv1 - AG2 - G3 Bv2 = 0 -G2 v1 + AG2 + G3 Bv2 = iA or in matrix form B AG1 + G2 B -G2 v1 0 -G2 - G3 B RB R = B R iA v2 AG2 + G3 B Note that the presence of the dependent source has destroyed the symmetrical nature of the node equations. v1 v2 Figure 3.8 Circuit with a dependent source. R2 R1 io R3 iA io Let us determine the node voltages for the network in Fig. 3.8, given the following parameters: = 2 3.3 iA = 2 mA R 2 = 6 k R 1 = 12 k EXAMPLE R 3 = 3 k Using these values with the equations for the network yields SOLUTION 1 1 V + V = 0 4k 1 2k 2 - 1 1 V + V = 2 * 10-3 6k 1 2k 2 Solving these equations using any convenient method yields V1 = -24兾5 V and V2 = 12兾5 V. We can check these answers by determining the branch currents in the network and then using that information to test KCL at the nodes. For example, the current from top to bottom through R 3 is Io = V2 12兾5 4 = = A R3 3k 5k Similarly, the current from right to left through R 2 is I2 = 12兾5 - (-24兾5) V2 - V1 6 = = A R2 6k 5k All the results are shown in Fig. 3.9. Note that KCL is satisfied at every node. –24 V1=— V 5 6 I2=— A 5k 12 V2=— V Figure 3.9 5 Circuit used in Example 3.3. 6 k⍀ 8 2Io=— A 5k 12 k⍀ 10 —A 5k 3 k⍀ –2 I1=— A 5k 4 Io=— A 5k irwin03_101-155hr.qxd 112 30-06-2010 CHAPTER 3 13:12 Page 112 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S EXAMPLE 3.4 Let us determine the set of linearly independent equations that when solved will yield the node voltages in the network in Fig. 3.10. Then given the following component values, we will compute the node voltages using MATLAB: R1 = 1 k, R2 = R3 = 2 k, R4 = 4 k, iA = 2 mA, iB = 4 mA, and = 2. Figure 3.10 iA Circuit containing a voltage-controlled current source. v1 v2 R1 + vx - v3 R2 ␣vx R3 R4 iB Applying KCL at each of the nonreference nodes yields the equations G3 v1 + G1 Av1 - v2 B - iA = 0 iA + G1 Av2 - v1 B + vx + G2 Av2 - v3 B = 0 G2 Av3 - v2 B + G4 v3 - iB = 0 where vx = v2 - v3 . Simplifying these equations, we obtain AG1 + G3 Bv1 - G1 v2 = iA -G1 v1 + AG1 + + G2 Bv2 - A + G2 Bv3 = -iA -G2 v2 + AG2 + G4 Bv3 = iB Given the component values, the equations become 1 1 + 1k 2k 1 F k 0 - 1 k 1 1 + 2 + k 2k 1 2k 0 v1 0.002 1 b V C v2 S = C -0.002 S 2k v3 0.004 1 1 + 2k 4k - a2 + or 0.0015 C -0.001 0 -0.001 0 v1 0.002 2.0015 -2.0005 S C v2 S = C -0.002 S -0.0005 0.00075 v3 0.004 Once again, the circuit equations resulting from a nodal analysis or in the form Gv = i, and the results obtained from perfoming the operation v = G-1 i are v1 = 11.9940 V v2 = 15.9910 V v3 = 15.9940 V irwin03_101-155hr.qxd 30-06-2010 13:12 Page 113 SECTION 3.1 N O D A L A N A LY S I S 113 Learning Assessments E3.4 Find the node voltages in the circuit in Fig. E3.4. V1 ANSWER: V1 = 16 V, V2 = -8 V. V2 10 k⍀ 2Io 10 k⍀ 10 k⍀ 4 mA Io Figure E3.4 E3.5 Find the voltage Vo in the network in Fig. E3.5. ANSWER: Vo = 4 V. Vx 2 mA 3 k⍀ + Vx — 12 k⍀ 6000 12 k⍀ Vo - Figure E3.5 ANSWER: Vo = 0.952 V. E3.6 Find Vo in Fig. E3.6 using nodal analysis. 2 mA 6 k⍀ 2 k⍀ IA + 2IA 3 k⍀ 6 k⍀ 1 k⍀ Vo – Figure E3.6 C I R C U I T S CO N TA I N I N G I N D E P E N D E N T VO LTA G E S O U R C E S As is our practice, in our discussion of this topic we will proceed from the simplest case to more complicated cases. The simplest case is that in which an independent voltage source is connected to the reference node. The following example illustrates this case. Consider the circuit shown in Fig. 3.11a. Let us determine all node voltages and branch currents. This network has three nonreference nodes with labeled node voltages V1 , V2 , and V3 . Based on our previous discussions, we would assume that in order to find all the node voltages we would need to write a KCL equation at each of the nonreference nodes. The resulting three linearly independent simultaneous equations would produce the unknown node voltages. However, note that V1 and V3 are known quantities because an independent voltage source is connected directly between the nonreference node and each of these nodes. Therefore, V1 = 12 V and V3 = -6 V. Furthermore, note that the current through the 9-kΩ resistor is [12 - (-6)]兾9k = 2 mA from left to right. We do not know V2 or the current in the remaining resistors. However, since only one node voltage is unknown, a single-node equation will produce it. Applying KCL to this center node yields EXAMPLE SOLUTION 3.5 irwin03_101-155hr.qxd 114 30-06-2010 CHAPTER 3 13:12 Page 114 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S [hint] V2 - V3 V2 - V1 V2 - 0 + + = 0 12k 6k 12k Any time an independent voltage source is connected between the reference node and a nonreference node, the nonreference node voltage is known. V2 - (-6) V2 - 12 V2 + + = 0 12k 6k 12k or from which we obtain V2 = 3 V 2 Once all the node voltages are known, Ohm’s law can be used to find the branch currents shown in Fig. 3.11b. The diagram illustrates that KCL is satisfied at every node. Note that the presence of the voltage sources in this example has simplified the analysis, since two of the three linear independent equations are V1 = 12 V and V3 = -6 V. We will find that as a general rule, whenever voltage sources are present between nodes, the node voltage equations that describe the network will be simpler. Figure 3.11 2 —A k 9 k⍀ Circuit used in Example 3.5. V2 12 k⍀ V1 12 k⍀ 7 —A 8k V3 9 k⍀ 3 —V 2 +12 V –6 V 12 k⍀ 12 V ± – + 6 k⍀ 5 —A 8k 6V 12 V ± – 6 k⍀ 23 —A 8k (a) 12 k⍀ + 1 —A 4k (b) Learning Assessment E3.7 Use nodal analysis to find the current Io in the network in Fig. E3.7. ANSWER: Io = Vo 6 k⍀ 6V 3 mA. 4 6 k⍀ ± – 3 k⍀ ± – 3V Io Figure E3.7 E3.8 Find V0 in Fig. E3.8 using nodal analysis. ANSWER: V0 = 3.89 V. 2 mA 6 k⍀ 2 k⍀ + 6 k⍀ 8 mA 3 k⍀ 1 k⍀ 12 V Figure E3.8 ± – Vo – 6V 21 —A 8k irwin03_101-155hr.qxd 30-06-2010 13:12 Page 115 SECTION 3.1 N O D A L A N A LY S I S 115 Next let us consider the case in which an independent voltage source is connected between two nonreference nodes. EXAMPLE Suppose we wish to find the currents in the two resistors in the circuit of Fig. 3.12a. SOLUTION If we try to attack this problem in a brute-force manner, we immediately encounter a problem. Thus far, branch currents were either known source values or could be expressed as the branch voltage divided by the branch resistance. However, the branch current through the 6-V source is certainly not known and cannot be directly expressed using Ohm’s law. We can, of course, give this current a name and write the KCL equations at the two nonreference nodes in terms of this current. However, this approach is no panacea because this technique will result in two linearly independent simultaneous equations in terms of three unknowns—that is, the two node voltages and the current in the voltage source. To solve this dilemma, we recall that N-1 linearly independent equations are required to determine the N-1 nonreference node voltages in an N-node circuit. Since our network has three nodes, we need two linearly independent equations. Now note that if somehow one of the node voltages is known, we immediately know the other; that is, if V1 is known, then V2 = V1 - 6. If V2 is known, then V1 = V2 + 6. Therefore, the difference in potential between the two nodes is constrained by the voltage source and, hence, 3.6 V1 - V2 = 6 This constraint equation is one of the two linearly independent equations needed to determine the node voltages. Next consider the network in Fig. 3.12b, in which the 6-V source is completely enclosed within the dashed surface. The constraint equation governs this dashed portion of the network. The remaining equation is obtained by applying KCL to this dashed surface, which is commonly called a supernode. Recall that in Chapter 2 we demonstrated that KCL must hold for a surface, and this technique eliminates the problem of dealing with a current through a voltage source. KCL for the supernode is -6 * 10-3 + V1 V2 + + 4 * 10-3 = 0 6k 12k Solving these equations yields V1 = 10 V and V2 = 4 V and, hence, I1 = 5兾3 mA and I2 = 1兾3 mA. A quick check indicates that KCL is satisfied at every node. Note that applying KCL at the reference node yields the same equation as shown above. The student may feel that the application of KCL at the reference node saves one from having to deal with supernodes. Recall that we do not apply KCL at any node—even the reference node—that contains an independent voltage source. This idea can be illustrated with the circuit in the next example. V1 6V ±– V2 V1 4 mA 6 mA 6 k⍀ V2 6V 12 k⍀ 6 mA (a) ±– 6 k⍀ 12 k⍀ I1 I2 (b) 4 mA Figure 3.12 Circuits used in Example 3.6. irwin03_101-155hr.qxd 116 30-06-2010 CHAPTER 3 Page 116 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S EXAMPLE 3.7 13:12 SOLUTION Let us determine the current Io in the network in Fig. 3.13a. Examining the network, we note that node voltages V2 and V4 are known and the node voltages V1 and V3 are constrained by the equation V1 - V3 = 12 The network is redrawn in Fig. 3.13b. Figure 3.13 V3+12 V1 Example circuit with supernodes. ± – 2 k⍀ V2 12 V V3 1 k⍀ ± – 2 k⍀ 12 V 2 k⍀ V3 1 k⍀ + ± – 2 k⍀ V4 1 k⍀ 6V 2 k⍀ 12 V 6V 1 k⍀ + 2 k⍀ Io ± – 12 V Io (a) (b) Since we want to find the current Io , V1 (in the supernode containing V1 and V3) is written as V3 + 12. The KCL equation at the supernode is then V3 + 12 - (-6) 2k V3 - (-6) V3 + 12 - 12 V3 - 12 V3 + + + = 0 2k 1k 1k 2k + Solving the equation for V3 yields V3 = - 6 V 7 Io can then be computed immediately as 6 7 3 Io = = - mA 2k 7 - Learning Assessment E3.9 Use nodal analysis to find Io in the network in Fig. E3.9. V1 V2 12 V -± ANSWER: Io = 3.8 mA. V3 2 k⍀ 6V ± - 2 k⍀ 1 k⍀ 2 k⍀ Io Figure E3.9 V4 ± 4V irwin03_101-155hr.qxd 30-06-2010 13:12 Page 117 SECTION 3.1 N O D A L A N A LY S I S 117 ANSWER: Vo = 5.6 V. E3.10 Find Vo in Fig. E3.10 using nodal analysis. 2 mA 12 V 2 k⍀ –+ 8 mA 3 k⍀ + 6 k⍀ 1 k⍀ Vo – Figure E3.10 C I R C U I T S CO N TA I N I N G D E P E N D E N T VO LTA G E S O U R C E S As the following examples will indicate, networks containing dependent (controlled) sources are treated in the same manner as described earlier. EXAMPLE We wish to find Io in the network in Fig. 3.14. Since the dependent voltage source is connected between the node labeled V1 and the reference node, SOLUTION V1 = 2kIx KCL at the node labeled V2 is V2 - V1 V2 4 + = 0 2k k 1k where Ix = V2 1k Solving these equations yields V2 = 8 V and V1 = 16 V. Therefore, Io = V1 - V2 2k = 4 mA 2 k⍀ V1 Io 2 kIx ± – 2 k⍀ Figure 3.14 V2 Ix 4 mA Circuits used in Example 3.8. 1 k⍀ 3.8 irwin03_101-155hr.qxd 118 30-06-2010 CHAPTER 3 Page 118 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S EXAMPLE 3.9 13:12 SOLUTION Let us find the current Io in the network in Fig. 3.15. This circuit contains both an independent voltage source and a voltage-controlled voltage source. Note that V3 = 6 V , V2 = Vx , and a supernode exists between the nodes labeled V1 and V2 . Applying KCL to the supernode, we obtain V2 - V3 V1 - V3 V1 V2 + + + = 0 6k 12k 6k 12k where the constraint equation for the supernode is V1 - V2 = 2Vx The final equation is V3 = 6 Solving these equations, we find that V1 = 9 V 2 and, hence, Io = Figure 3.15 V1 3 = mA 12k 8 6 k⍀ Circuit used in Example 3.9. 2Vx ±– V1 12 k⍀ V2 V3 + 12 k⍀ 6 k⍀ Io Vx ± – 6V - Finally, let us consider two additional circuits that, for purposes of comparison, we will examine using more than one method. EXAMPLE 3.10 Let us find Vo in the network in Fig. 3.16a. Note that the circuit contains two voltage sources, one of which is a controlled source, and two independent current sources. The circuit is redrawn in Fig. 3.16b in order to label the nodes and identify the supernode surrounding the controlled source. Because of the presence of the independent voltage source, the voltage at node 4 is known to be 4 V. We will use this knowledge in writing the node equations for the network. irwin03_101-155hr.qxd 30-06-2010 13:12 Page 119 SECTION 3.1 119 N O D A L A N A LY S I S Since the network has five nodes, four linear independent equations are sufficient to determine all the node voltages. Within the supernode, the defining equation is V1 - V2 = 2Vx where V2 = Vx and thus V1 = 3Vx Furthermore, we know that one additional equation is V4 = 4 Thus, given these two equations, only two more equations are needed in order to solve for the unknown node voltages. These additional equations result from applying KCL at the supernode and at the node labeled V3 . The equations are - Vx - V3 3Vx - V3 Vx 3Vx - 4 2 + + + + = 0 k 1k 1k 1k 1k V3 - Vx V3 - 3Vx 2 + = 1k 1k k Combining the equations yields the two equations 8Vx - 2V3 = 6 -4Vx + 2V3 = 2 Solving these equations, we obtain Vx = 2 V and V3 = 5 V Vo = 3Vx - V3 = 1 V V1 ± – + 2Vx 1 k⍀ Vo - 2 —A k 2 mA + Vx ± – 1 k⍀ 1 k⍀ 1 k⍀ - V2 + 2 mA ± – 4V + Vo - 2Vx Vx 1 k⍀ V3 1 k⍀ V4 1 k⍀ 2 —A k 1 k⍀ - (a) ± – 4V (b) Figure 3.16 Circuit used in Example 3.10. We wish to find Io in the network in Fig. 3.17a. Note that this circuit contains three voltage sources, one of which is a controlled source and another is a controlled current source. Because two of the voltage sources are connected to the reference node, one node voltage is known directly and one is specified by the dependent source. Furthermore, the difference in voltage between two nodes is defined by the 6-V independent source. EXAMPLE 3.11 irwin03_101-155hr.qxd 120 30-06-2010 CHAPTER 3 13:12 Page 120 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S The network is redrawn in Fig. 3.17b in order to label the nodes and identify the supernode. Since the network has six nodes, five linear independent equations are needed to determine the unknown node voltages. The two equations for the supernode are V1 - V4 = -6 V4 - V3 V1 - V3 V4 - V5 V4 V1 - 12 + + 2Ix + + + = 0 1k 1k 1k 1k 1k The three remaining equations are V2 = 12 V3 = 2Vx V5 V5 - V4 + = 2Ix 1k 1k The equations for the control parameters are Vx = V1 - 12 Ix = V4 1k Combining these equations yields the following set of equations: -2V1 + 5V4 - V5 = -36 V1 - V4 = -6 -3V4 + 2V5 = 0 Solving these equations by any convenient means yields V1 = -38 V V4 = -32 V V5 = -48 V Then, since V3 = 2Vx , V3 = -100 V. Io is -48 mA. The reader is encouraged to verify that KCL is satisfied at every node. V1 + Vx 1 k⍀ - + 1 k⍀ 1 k⍀ 6V 1 k⍀ 2Ix + Vx V2 1 k⍀ 1 k⍀ V3 ± – ± – 1 k⍀ 1 k⍀ 2Vx Ix (a) Figure 3.17 Circuit used in Example 3.11. 6V 1 k⍀ Io 12 V ± – ± – 2Ix V4 1 k⍀ 12 V + 1 k⍀ 2Vx (b) V5 1 k⍀ 1 k⍀ 1 k⍀ Ix Io irwin03_101-155hr.qxd 30-06-2010 13:12 Page 121 SECTION 3.1 N O D A L A N A LY S I S 121 Problem-Solving Strategy Step 1. Determine the number of nodes in the circuit. Select one node as the reference Nodal Analysis node. Assign a node voltage between each nonreference node and the reference node. All node voltages are assumed positive with respect to the reference node. For an N-node circuit, there are N - 1 node voltages. As a result, N - 1 linearly independent equations must be written to solve for the node voltages. Step 2. Write a constraint equation for each voltage source—independent or dependent— in the circuit in terms of the assigned node voltages using KVL. Each constraint equation represents one of the necessary linearly independent equations, and Nv voltage sources yield Nv linearly independent equations. For each dependent voltage source, express the controlling variable for that source in terms of the node voltages. A voltage source—independent or dependent—may be connected between a nonreference node and the reference node or between two nonreference nodes. A supernode is formed by a voltage source and its two connecting nonreference nodes. Step 3. Use KCL to formulate the remaining N - 1 - Nv linearly independent equa- tions. First, apply KCL at each nonreference node not connected to a voltage source. Second, apply KCL at each supernode. Treat dependent current sources like independent current sources when formulating the KCL equations. For each dependent current source, express the controlling variable in terms of the node voltages. Learning Assessment E3.11 Use nodal analysis to find Io in the circuit in Fig. E3.11. V1 4 mA 2000 Ix + - ANSWER: Io = 4 mA. 3 V2 2 k⍀ 2 mA 2 k⍀ Ix Io Figure E3.11 E3.12 Find Vo in Fig. E3.12 using nodal analysis. ANSWER: Vo = 6.29 V. 2 mA 4kIx +– 8 mA 3 k⍀ 2 k⍀ Ix + 6 k⍀ 1 k⍀ Vo – Figure E3.12 irwin03_101-155hr.qxd 122 30-06-2010 CHAPTER 3 13:12 Page 122 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.2 Loop Analysis We found that in a nodal analysis the unknown parameters are the node voltages and KCL was employed to determine them. Once these node voltages have been calculated, all the branch currents in the network can easily be determined using Ohm’s law. In contrast to this approach, a loop analysis uses KVL to determine a set of loop currents in the circuit. Once these loop currents are known, Ohm’s law can be used to calculate any voltages in the network. Via network topology we can show that, in general, there are exactly B - N + 1 linearly independent KVL equations for any network, where B is the number of branches in the circuit and N is the number of nodes. For example, if we once again examine the circuit in Fig. 2.5, we find that there are eight branches and five nodes. Thus, the number of linearly independent KVL equations necessary to determine all currents in the network is B - N + 1 = 8 - 5 + 1 = 4. The network in Fig. 2.5 is redrawn as shown in Fig. 3.18 with 4 loop currents labeled as shown. The branch currents are then determined as Figure 3.18 1 Figure 2.5 redrawn with loop currents. i2(t) i3(t) R2 R1 iB(t) v (t) i5(t) i4(t) 3 1 4 ±– iA(t) i1(t) 2 R3 R4 iC(t) ± – i6(t) v2(t) iD(t) i7(t) R5 i8(t) 5 i1(t) = iA(t) i2(t) = iA(t) - iB(t) i3(t) = iB(t) i4(t) = iA(t) - iC(t) i5(t) = iB(t) - iD(t) i6(t) = -iC(t) i7(t) = iC(t) - iD(t) i8(t) = -iD(t) All the circuits we will examine in this text will be planar, which simply means that we can draw the circuit on a sheet of paper in such a way that no conductor crosses another conductor. If a circuit is planar, the loops are more easily identified. For example, recall in Chapter 2 that we found that a single equation was sufficient to determine the current in a circuit containing a single loop. If the circuit contains N independent loops, we will show (and the general topological formula B - N + 1 can be used for verification) that N independent simultaneous equations will be required to describe the network. Our approach to loop analysis will mirror the approach used in nodal analysis (i.e., we will begin with simple cases and systematically proceed to those that are more difficult). Then at the end of this section we will outline a general strategy for employing loop analysis. C I R C U I T S CO N TA I N I N G O N LY I N D E P E N D E N T VO LTA G E S O U R C E S To begin our analysis, consider the circuit shown in Fig. 3.19. We note that this network has seven branches and six nodes, and thus the number of linearly independent KVL equations necessary to determine all currents in the circuit is B - N + 1 = 7 - 6 + 1 = 2. Since two linearly independent KVL equations are required, we identify two independent loops, A-B-E-F-A and B-C-D-E-B. We now define a new set of current variables called loop currents, which can be used to find the physical currents in the circuit. Let us assume that current i1 flows in the first loop and that current i2 flows in the second loop. Then the branch current flowing from B to E irwin03_101-155hr.qxd 30-06-2010 13:12 Page 123 SECTION 3.2 A v + 1R1 vS1 ± – - C ±– + R3 v3 i1 - R2 F vS2 B v2 + E - L O O P A N A LY S I S 123 Figure 3.19 A two-loop circuit. + i2 R4 v4 R5 - v5 + D through R3 is i1 - i2. The directions of the currents have been assumed. As was the case in the nodal analysis, if the actual currents are not in the direction indicated, the values calculated will be negative. Applying KVL to the first loop yields +v1 + v3 + v2 - vS1 = 0 KVL applied to loop 2 yields [hint] The equations employ the passive sign convention. +vS2 + v4 + v5 - v3 = 0 where v1 = i1R1 , v2 = i1R2 , v3 = Ai1 - i2 B R3 , v4 = i2R4 , and v5 = i2R5 . Substituting these values into the two KVL equations produces the two simultaneous equations required to determine the two loop currents; that is, i1 AR1 + R2 + R3 B - i2 AR3 B = vS1 -i1 AR3 B + i2 AR3 + R4 + R5 B = -vS2 or in matrix form B R1 + R2 + R3 -R3 -R3 i v R B 1 R = B S1 R R3 + R4 + R5 i2 -vS2 At this point, it is important to define what is called a mesh. A mesh is a special kind of loop that does not contain any loops within it. Therefore, as we traverse the path of a mesh, we do not encircle any circuit elements. For example, the network in Fig. 3.19 contains two meshes defined by the paths A-B-E-F-A and B-C-D-E-B. The path A-B-C-D-E-F-A is a loop, but it is not a mesh. Since the majority of our analysis in this section will involve writing KVL equations for meshes, we will refer to the currents as mesh currents and the analysis as a mesh analysis. Consider the network in Fig. 3.20a. We wish to find the current Io . We will begin the analysis by writing mesh equations. Note that there are no + and - signs on the resistors. However, they are not needed, since we will apply Ohm’s law to each resistive element as we write the KVL equations. The equation for the first mesh is -12 + 6kI1 + 6kAI1 - I2 B = 0 The KVL equation for the second mesh is 6kAI2 - I1 B + 3kI2 + 3 = 0 where Io = I1 - I2 . Solving the two simultaneous equations yields I1 = 5兾4 mA and I2 = 1兾2 mA. Therefore, Io = 3兾4 mA. All the voltages and currents in the network are shown in Fig. 3.20b. Recall from nodal analysis that once the node voltages were determined, we could check our analysis using KCL at the nodes. In this case, we know the branch currents and can use KVL around any closed path to check our results. For example, applying KVL to the outer loop yields -12 + 3 15 + + 3 = 0 2 2 0 = 0 EXAMPLE SOLUTION 3.12 irwin03_101-155hr.qxd 124 30-06-2010 CHAPTER 3 13:12 Page 124 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S Since we want to calculate the current Io , we could use loop analysis, as shown in Fig. 3.20c. Note that the loop current I1 passes through the center leg of the network and, therefore, I1 = Io . The two loop equations in this case are -12 + 6kAI1 + I2 B + 6kI1 = 0 and -12 + 6kAI1 + I2 B + 3kI2 + 3 = 0 Solving these equations yields I1 = 3兾4 mA and I2 = 1兾2 mA. Since the current in the 12-V source is I1 + I2 = 5兾4 mA, these results agree with the mesh analysis. Finally, for purposes of comparison, let us find Io using nodal analysis. The presence of the two voltage sources would indicate that this is a viable approach. Applying KCL at the top center node, we obtain Vo - 12 Vo Vo - 3 + + = 0 6k 6k 3k and hence, Vo = 9 V 2 and then Io = Vo 3 = mA 6k 4 Note that in this case we had to solve only one equation instead of two. Figure 3.20 15 —V + 2 - Vo Circuits used in Example 3.12. 6 k⍀ 12 V ± – 5 — mA 4 3 k⍀ 6 k⍀ I1 ± – I2 3V 3 —V + 2 - Io 3 k⍀ + 9 — mA 6 k⍀ 2 3 — mA 4 1 — mA 2 6 k⍀ ± – 12 V Vo ± – 3V (b) (a) 6 k⍀ 12 V ± – Vo I1 3 k⍀ 6 k⍀ I2 ± – 3V Io (c) Once again we are compelled to note the symmetrical form of the mesh equations that describe the circuit in Fig. 3.19. Note that the coefficient matrix for this circuit is symmetrical. Since this symmetry is generally exhibited by networks containing resistors and independent voltage sources, we can learn to write the mesh equations by inspection. In the first equation, the coefficient of i1 is the sum of the resistances through which mesh current 1 flows, and the coefficient of i2 is the negative of the sum of the resistances common to mesh current 1 and mesh current 2. The right-hand side of the equation is the algebraic sum of the voltage sources in mesh 1. The sign of the voltage source is positive if it aids the assumed direction of the current flow and negative if it opposes the assumed flow. The first equation is KVL for mesh 1. In the second equation, the coefficient of i2 is the sum of all the irwin03_101-155hr.qxd 30-06-2010 13:12 Page 125 SECTION 3.2 L O O P A N A LY S I S 125 resistances in mesh 2, the coefficient of i1 is the negative of the sum of the resistances common to mesh 1 and mesh 2, and the right-hand side of the equation is the algebraic sum of the voltage sources in mesh 2. In general, if we assume all of the mesh currents to be in the same direction (clockwise or counterclockwise), then if KVL is applied to mesh j with mesh current ij , the coefficient of ij is the sum of the resistances in mesh j and the coefficients of the other mesh currents Ae.g., ij - 1 , ij + 1 B are the negatives of the resistances common to these meshes and mesh j. The right-hand side of the equation is equal to the algebraic sum of the voltage sources in mesh j. These voltage sources have a positive sign if they aid the current flow ij and a negative sign if they oppose it. Let us write the mesh equations by inspection for the network in Fig. 3.21. Then we will use MATLAB to solve for the mesh currents. EXAMPLE SOLUTION The three linearly independent simultaneous equations are 3.13 (4k + 6k)I1 - (0)I2 - (6k)I3 = -6 -(0)I1 + (9k + 3k)I2 - (3k)I3 = 6 -(6k)I1 - (3k)I2 + (3k + 6k + 12k)I3 = 0 or in matrix form 10k C 0 -6k 0 12k -3k -6k I1 -6 -3k S C I2 S = C 6 S 21k I3 0 Note the symmetrical form of the equations. The general form of the matrix equation is RI = V and the solution of this matrix equation is I = R - 1V Performing the indicated operation yields the following loop currents: i1 = -0.6757 mA i2 = 0.4685 mA i3 = -0.1261 mA Figure 3.21 4 k⍀ I1 + Circuit used in Example 3.13. 6 k⍀ 6V 9 k⍀ I2 I3 3 k⍀ 12 k⍀ irwin03_101-155hr.qxd 126 30-06-2010 CHAPTER 3 13:12 Page 126 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S C I R C U I T S CO N TA I N I N G I N D E P E N D E N T C U R R E N T S O U R C E S Just as the presence of a voltage source in a network simplified the nodal analysis, the presence of a current source simplifies a loop analysis. The following examples illustrate the point. Learning Assessment E3.13 Use mesh equations to find Vo in the circuit in Fig. E3.13. ANSWER: Vo = 33 V. 5 6V + + 4 k⍀ 2 k⍀ 2 k⍀ Vo 6 k⍀ ± – 3V - Figure E3.13 E3.14 Find Vo in Fig. E3.14 using mesh analysis. ANSWER: Vo = 8.96 V. 3 k⍀ 2 k⍀ 4 k⍀ + 4 k⍀ 12 V 6 k⍀ + 3 k⍀ 10 V + – Figure E3.14 EXAMPLE 3.14 SOLUTION 8V + – 6 k⍀ Vo – Let us find both Vo and V1 in the circuit in Fig. 3.22. Although it appears that there are two unknown mesh currents, the current I1 goes directly through the current source and, therefore, I1 is constrained to be 2 mA. Hence, only the current I2 is unknown. KVL for the rightmost mesh is 2k(I2 - I1) - 2 + 6kI2 = 0 And, of course, I1 = 2 * 10-3 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 127 SECTION 3.2 L O O P A N A LY S I S 127 These equations can be written as - 2kI1 + 8kI2 = 2 I1 = 2兾k Solving these equation for I2 yields I2 = 3/4kA and thus Vo = 6kI2 = 9 V 2 To obtain V1 we apply KVL around any closed path. If we use the outer loop, the KVL equation is -V1 + 4kI1 - 2 + 6kI2 = 0 And therefore, V1 = 21 V 2 Note that since the current I1 is known, the 4-k resistor did not enter the equation in finding Vo . However, it appears in every loop containing the current source and, thus, is used in finding V1 . 2V V1 4 k⍀ 2 mA Figure 3.22 + I1 I2 + 6 k⍀ 2 k⍀ Vo - We wish to find Vo in the network in Fig. 3.23. EXAMPLE Since the currents I1 and I2 pass directly through a current source, two of the three required equations are I1 = 4 * 10-3 I2 = -2 * 10-3 The third equation is KVL for the mesh containing the voltage source; that is, 4kAI3 - I2 B + 2kAI3 - I1 B + 6kI3 - 3 = 0 These equations yield I3 = 1 mA 4 and hence, Vo = 6kI3 - 3 = Circuit used in Example 3.14. -3 V 2 SOLUTION 3.15 irwin03_101-155hr.qxd 128 30-06-2010 CHAPTER 3 13:12 Page 128 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S Figure 3.23 4 mA Circuit used in Example 3.15. + I1 2 k⍀ 6 k⍀ I3 Vo 4 k⍀ 2 mA + 4 k⍀ I2 3V - What we have demonstrated in the previous example is the general approach for dealing with independent current sources when writing KVL equations; that is, use one loop through each current source. The number of “window panes” in the network tells us how many equations we need. Additional KVL equations are written to cover the remaining circuit elements in the network. The following example illustrates this approach. EXAMPLE 3.16 SOLUTION [hint] In this case the 4-mA current source is located on the boundary between two meshes. Thus, we will demonstrate two techniques for dealing with this type of situtation. One is a special loop technique, and the other is known as the supermesh approach. Let us find Io in the network in Fig. 3.24a. First, we select two loop currents I1 and I2 such that I1 passes directly through the 2-mA source, and I2 passes directly through the 4-mA source, as shown in Fig. 3.24b. Therefore, two of our three linearly independent equations are I1 = 2 * 10-3 I2 = 4 * 10-3 The remaining loop current I3 must pass through the circuit elements not covered by the two previous equations and cannot, of course, pass through the current sources. The path for this remaining loop current can be obtained by open-circuiting the current sources, as shown in Fig. 3.24c. When all currents are labeled on the original circuit, the KVL equation for this last loop, as shown in Fig. 3.24d, is -6 + 1kI3 + 2kAI2 + I3 B + 2kAI3 + I2 - I1 B + 1kAI3 - I1 B = 0 Solving the equations yields I3 = -2 mA 3 and therefore, Io = I1 - I2 - I3 = -4 mA 3 Next consider the supermesh technique. In this case the three mesh currents are specified as shown in Fig. 3.24e, and since the voltage across the 4-mA current source is unknown, it is assumed to be Vx . The mesh currents constrained by the current sources are I1 = 2 * 10-3 I2 - I3 = 4 * 10-3 The KVL equations for meshes 2 and 3, respectively, are 2kI2 + 2kAI2 - I1 B - Vx = 0 -6 + 1kI3 + Vx + 1kAI3 - I1 B = 0 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 129 SECTION 3.2 6V 1 k⍀ + 4 mA 1 k⍀ 2 k⍀ 2 mA 6V 4 mA I1 I2 2 k⍀ 2 mA Io (a) 1 k⍀ + I3 1 k⍀ 2 k⍀ 6V 1 k⍀ + I3 1 k⍀ 2 k⍀ I1 2 k⍀ 2 mA 6V 1 k⍀ + - 2 k⍀ I1 I2 I3 1 k⍀ 2 k⍀ I1 2 k⍀ Io (e) 1 k⍀ + Vx + 4 mA 2 k⍀ I2 (d) 6V I3 4 mA Io (c) 1 k⍀ 2 k⍀ Io (b) 6V 2 mA Circuits used in Example 3.16. 1 k⍀ 2 k⍀ Figure 3.24 1 k⍀ + 2 mA L O O P A N A LY S I S I2 2 k⍀ Io (f) Adding the last two equations yields -6 + 1kI3 + 2kI2 + 2kAI2 - I1 B + 1kAI3 - I1 B = 0 Note that the unknown voltage Vx has been eliminated. The two constraint equations, together with this latter equation, yield the desired result. The purpose of the supermesh approach is to avoid introducing the unknown voltage Vx . The supermesh is created by mentally removing the 4-mA current source, as shown in Fig. 3.24f. Then writing the KVL equation around the dotted path, which defines the supermesh, using the original mesh currents as shown in Fig. 3.24e, yields -6 + 1kI3 + 2kI2 + 2kAI2 - I1 B + 1kAI3 - I1 B = 0 Note that this supermesh equation is the same as that obtained earlier by introducing the voltage Vx . 129 irwin03_101-155hr.qxd 130 30-06-2010 CHAPTER 3 13:12 Page 130 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S Learning Assessments E3.15 Find Vo in the network in Fig. E3.15. ANSWER: Vo = 33 V. 5 ANSWER: Vo = 32 V. 5 + Vo 4 mA 6 k⍀ 2 k⍀ ± – 4 k⍀ 5V Figure E3.15 E3.16 Find Vo in the network in Fig. E3.16. 4 mA ± – 2 mA + 1 k⍀ 2 k⍀ 4 k⍀ 4V Vo - Figure E3.16 E3.17 Find Vo in Fig. E3.17 using loop analysis. ANSWER: Vo = 9.71 V. 3 k⍀ 2 k⍀ 4 k⍀ + 2 mA 12 V 6 k⍀ + 3 k⍀ 10 V + – Figure E3.17 E3.18 Find Vo in Fig. E3.17 using mesh analysis. 8V + – 6 k⍀ Vo – ANSWER: Vo = 9.71 V. irwin03_101-155hr.qxd 30-06-2010 13:12 Page 131 SECTION 3.2 L O O P A N A LY S I S 131 C I R C U I T S CO N TA I N I N G D E P E N D E N T S O U R C E S We deal with circuits containing dependent sources just as we have in the past. First, we treat the dependent source as though it were an independent source when writing the KVL equations. Then we write the controlling equation for the dependent source. The following examples illustrate the point. EXAMPLE Let us find Vo in the circuit in Fig. 3.25, which contains a voltage-controlled voltage source. SOLUTION The equations for the loop currents shown in the figure are - 2Vx + 2k(I1 + I2) + 4kI1 = 0 - 2Vx + 2k(I1 + I2) - 3 + 6kI2 = 0 where Vx = 4kI1 These equations can be combined to produce - 2kI1 + 2kI2 = 0 - 6kI1 + 8kI2 = 3 These equations can be placed in the form RI = V, where -2000 2000 i1 R= 0 I= -6000 8000 V= i2 3 The solution is I = R-1V, and this operation yields i1 = 1.5 mA i2 = 1.5 mA and therefore, Vo = 6kI2 = 9 V For comparison, we will also solve the problem using nodal analysis. The presence of the voltage sources indicates that this method could be simpler. Treating the 3-V source and its connecting nodes as a supernode and writing the KCL equation for this supernode yields Vx - 2Vx Vx Vx + 3 + + = 0 2k 4k 6k where Vo = Vx + 3 These equations also yield Vo = 9 V. Vx 2 k⍀ 2 Vx ± – I1 3V Figure 3.25 + 4 k⍀ + 6 k⍀ I2 Vo - Circuit used in Example 3.17. 3.17 irwin03_101-155hr.qxd 132 30-06-2010 CHAPTER 3 13:12 Page 132 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S EXAMPLE 3.18 SOLUTION Let us find Vo in the circuit in Fig. 3.26, which contains a voltage-controlled current source. The currents I1 and I2 are drawn through the current sources. Therefore, two of the equations needed are Vx 2000 I2 = 2 * 10 - 3 I1 = The KVL equation for the third mesh is - 3 + 2k(I3 - I1) + 6kI3 = 0 where Vx = 4k (I1 - I2) Combining these equations yields - I1 + 2I2 = 0 I2 = 2兾k - 2kI2 + 8kI3 = 3 The equations can be expressed in matrix form as IR = V, where R= -1 0 -2000 2 1 0 0 0 8000 i1 I = i2 i3 0 and V = 0.002 3 Performing the operation I = R-1V, produces the currents i1 = 4.0 mA i2 = 2.0 mA i3 = 1.375 mA And hence, Vo = 8.25 V. Figure 3.26 Circuit used in Example 3.18. + Vx — 2000 I1 2 k⍀ -Vx + I3 6 k⍀ 4 k⍀ 2 mA I2 ± – Vo 3V - EXAMPLE 3.19 SOLUTION The network in Fig. 3.27 contains both a current-controlled voltage source and a voltagecontrolled current source. Let us use MATLAB to determine the loop currents. The equations for the loop currents shown in the figure are 4 k Vx I2 = 2k -1kIx + 2kAI3 - I1 B + 1kAI3 - I4 B = 0 I1 = 1kAI4 - I3 B + 1kAI4 - I2 B + 12 = 0 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 133 SECTION 3.2 where L O O P A N A LY S I S 133 Vx = 2kAI3 - I1 B Ix = I4 - I2 Combining these equations yields 4 k I1 + I2 - I3 = 0 I1 = 1kI2 + 3kI3 - 2kI4 = 8 1kI2 + 1kI3 - 2kI4 = 12 In matrix form the equations are 1 1 D 0 0 0 1 1k 1k 4 0 I1 k 0 I T D 2T = E 0 U -2k I3 8 -2k I4 12 0 -1 3k 1k The equations are in the form RI = V, and the solution to I = R-1V is i1 = 4.0 mA i2 = 6.0 mA i3 = -2.0 mA i4 = -1.0 mA Figure 3.27 2 k⍀ I1 4 mA Vx I2 — 2k Circuit used in Example 3.19. + Vx - Ix 1kIx ± – 2 k⍀ 1 k⍀ I3 I4 ± – 12 V 1 k⍀ At this point we will again examine the circuit in Example 3.10 and analyze it using loop equations. Recall that because the network has two voltage sources, the nodal analysis was somewhat simplified. In a similar manner, the presence of the current sources should simplify a loop analysis. Clearly, the network has four loops, and thus four linearly independent equations are required to determine the loop currents. The network is redrawn in Fig. 3.28 where the loop currents are specified. Note that we have drawn one current through each of the independent current sources. This choice of currents simplifies the analysis since two of the four equations are I1 = 2兾k I3 = -2兾k EXAMPLE 3.20 irwin03_101-155hr.qxd 134 30-06-2010 CHAPTER 3 13:12 Page 134 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S The two remaining KVL equations for loop currents I2 and I4 are -2Vx + 1kI2 + (I2 - I3)1k = 0 (I4 + I3 - I1)1k - 2Vx + 1kI4 + 4 = 0 where Vx = 1k(I1 - I3 - I4) Substituting the equations for I1 and I3 into the two KVL equations yields 2kI2 + 2kI4 = 6 4kI4 = 8 Solving these equations for I2 and I4, we obtain I4 = 2 mA I2 = 1 mA and thus Vo = 1V Figure 3.28 + Circuit used in Example 3.20. ± – 2Vx 2 —A k I1 I2 Vo - + - 3.21 1 k⍀ 1 k⍀ I4 1 k⍀ Vx EXAMPLE 1 k⍀ 2 —A k I3 ± – 4V Let us once again consider Example 3.11. In this case we will examine the network using loop analysis. Although there are four sources, two of which are dependent, only one of them is a current source. Thus, from the outset we expect that a loop analysis will be more difficult than a nodal analysis. Clearly, the circuit contains six loops. Thus, six linearly independent equations are needed to solve for all the unknown currents. The network is redrawn in Fig. 3.29 where the loops are specified. The six KVL equations that describe the network are 1kI1 + 1kAI1 - I2 B + 1kAI1 - I4 B = 0 1kAI2 - I1 B - 6 + 1kAI2 - I5 B = 0 I3 = 2Ix Figure 3.29 Circuit used in Example 3.21. + Vx 1 k⍀ 1 k⍀ 6V I1 I2 1 k⍀ 1 k⍀ + 2Ix I3 - 12 V ± – I4 2Vx ± – 1 k⍀ 1 k⍀ I5 I6 Ix 1 k⍀ Io irwin03_101-155hr.qxd 30-06-2010 13:12 Page 135 SECTION 3.2 -12 + 1kAI4 - I1 B + 2Vx = 0 -2Vx + 1kAI5 - I2 B + 1kAI5 - Io B = 0 1kAIo - I5 B + 1kAIo - I3 B + 1kIo = 0 And the control variables for the two dependent sources are Vx = -1kI1 Ix = I5 - Io Substituting the control parameters into the six KVL equations yields 3I1 -I1 0 -3I1 2I1 0 0 -I2 0 -I4 +2I2 0 0 -I5 0 I3 0 -2I5 0 0 +I4 0 -I2 0 0 +2I5 0 0 0 -3I5 0 0 +2Io 0 -Io +5Io = = = = = = 0 6兾k 0 12兾k 0 0 which can be written in matrix form as 3 -1 0 F -3 2 0 -1 2 0 0 -1 0 0 0 1 0 0 0 -1 0 0 1 0 0 0 -1 -2 0 2 -3 0 I1 0 0 I2 6兾k 2 I3 0 VF V = F V 0 12兾k I4 -1 I5 0 5 Io 0 The solution to the matrix equations RI = V is i1 = 50.0 mA i2 = -12.0 mA i3 = -64.0 mA i4 = 162.0 mA i5 = -80.0 mA i6 = -48.0 mA As a final point, it is very important to examine the circuit carefully before selecting an analysis approach. One method could be much simpler than another, and a little time invested up front may save a lot of time in the long run. For an N-node circuit, N - 1 linearly independent equations must be formulated to solve for N - 1 node voltages. An N-loop circuit requires the formulation of N linearly independent equations. One consideration in the selection of a method should be the number of linearly independent equations that must be formulated. The same circuit was solved in Example 3.10 using nodal analysis and in Example 3.20 using loop analysis. The circuit in Fig. 3.16 has four unknown node voltages. As a result, four linearly independent equations are required. Because there are two voltage sources, two constraint equations are needed. It was pointed out in Example 3.20 that this same circuit has four loops which requires four linearly independent equations. The two current sources produce two constraint equations. The effort required to solve this circuit using either nodal or loop analysis is similar. However, this is not true for many circuits. Consider the circuit in Fig. 3.30. This circuit has eight loops. Selection of the loop currents such that only one loop current flows through the independent current source leaves us with seven unknown loop currents. Since this circuit has seven nodes, there are six node voltages, and we must formulate six linearly independent L O O P A N A LY S I S 135 irwin03_101-155hr.qxd 136 30-06-2010 CHAPTER 3 13:12 Page 136 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S Figure 3.30 A circuit utilized in a discussion of the selection of an analysis technique. equations. By judicious selection of the bottom node as the reference node, four of the node voltages are known, leaving just two unknown node voltages—the node voltage across the current source and the node voltage across the 3- and 6- resistors. Applying KCL at these two nodes yields two equations that can be solved for the two unknown node voltages. Even with the use of a modern calculator or a computer program such as MATLAB, the solution of two simultaneous equations requires less effort than the solution of the seven simultaneous equations that the loop analysis would require. 2⍀ 20 V ± – 2⍀ 4⍀ 4⍀ 2A 16 V 2⍀ ± – 12 V 2⍀ ± – 2⍀ 3⍀ 6⍀ ± – 20 V Problem-Solving Strategy Loop Analysis Step 1. Determine the number of independent loops in the circuit. Assign a loop cur- rent to each independent loop. For an N-loop circuit, there are N-loop currents. As a result, N linearly independent equations must be written to solve for the loop currents. If current sources are present in the circuit, either of two techniques can be employed. In the first case, one loop current is selected to pass through one of the current sources. The remaining loop currents are determined by opencircuiting the current sources in the circuit and using this modified circuit to select them. In the second case, a current is assigned to each mesh in the circuit. Step 2. Write a constraint equation for each current source—independent or dependent— in the circuit in terms of the assigned loop current using KCL. Each constraint equation represents one of the necessary linearly independent equations, and NI current sources yield NI linearly independent equations. For each dependent current source, express the controlling variable for that source in terms of the loop currents. Step 3. Use KVL to formulate the remaining N - NI linearly independent equations. Treat dependent voltage sources like independent voltage sources when formulating the KVL equations. For each dependent voltage source, express the controlling variable in terms of the loop currents. Learning Assessments E3.19 Use mesh analysis to find Vo in the circuit in Fig. E3.19. 2 k⍀ 2000Ix Figure E3.19 ± – ANSWER: Vo = 12 V. 12 V + 4 k⍀ + Ix 2 k⍀ Vo - irwin03_101-155hr.qxd 30-06-2010 13:12 Page 137 SECTION 3.3 A P P L I C AT I O N E X A M P L E 137 E3.20 Use loop analysis to solve the network in Example 3.5 and compare the time and effort involved in the two solution techniques. E3.21 Use nodal analysis to solve the circuit in Example 3.15 and compare the time and effort involved in the two solution strategies. ANSWER: Vo = 6.97 V. E3.22 Find Vo in Fig. E3.22 using mesh analysis. 3 k⍀ 4 k⍀ + 4 k⍀ 2 k⍀ 12 V 6 k⍀ – 3 k⍀ 10 V + – 0.5Vx Figure E3.22 Vx + + Vo 6 k⍀ + – – ANSWER: Vo = 9 V. E3.23 Find Vo in Fig. E3.23 using mesh analysis. 3 k⍀ 2 k⍀ 4 k⍀ + 0.5Ix 12 V 6 k⍀ + 3 k⍀ 10 V Figure E3.23 + – 8V + – 6 k⍀ Ix Vo – 3.3 Application Example A conceptual circuit for manually setting the speed of a dc electric motor is shown in Fig. 3.31a. The resistors R1 and R2 are inside a component called a potentiometer, or pot, which is nothing more than an adjustable resistor, for example, a volume control. Turning the knob changes the ratio a = R2兾(R1 + R2), but the total resistance, Rpot = R1 + R2 , is APPLICATION EXAMPLE 3.22 • irwin03_101-155hr.qxd 138 30-06-2010 CHAPTER 3 13:12 Page 138 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S unchanged. In this way the pot forms a voltage divider that sets the voltage Vspeed . The power amplifier output, VM , is four times Vspeed . Power amplifiers can output the high currents needed to drive the motor. Finally, the dc motor speed is proportional to VM ; that is, the speed in rpm is some constant k times V. Without knowing the details of the power amplifier, can we analyze this system? In particular, can we develop a relationship between rpm and a? Figure 3.31 (a) A simple dc motor driver and (b) the circuit model used to analyze it. 5V R1 Rpot µ R2 ␣=1 ␣=0 + Vspeed - Power amp + VM/Vspeed=4 VM - dc motor (a) Power amp model 5V R1 Rpot µ R2 ␣=1 ␣=0 + Vspeed - + 4Vspeed - ± – + VM - (b) SOLUTION Since the power amplifier output voltage is proportional to its input, we can model the amplifier as a simple dependent source. The resulting circuit diagram is shown in Fig. 3.31b. Now we can easily develop a relationship between motor speed and the pot position, a. The equations that govern the operation of the motor, power amplifier, and the voltage divider are speed (rpm) = KM VM VM = 4Vspeed Vspeed = 5 R2 R2 = 5c d = 5a R1 + R2 Rpot R2 = aRpot R1 = (1 - a)Rpot Combining these relationships to eliminate Vspeed yields a relationship between motor speed and a, that is, rpm = 20a. If, for example, the motor constant KM is 50 rpm兾V, then rpm = 1000a This relationship specifies that the motor speed is proportional to the pot knob position. Since the maximum value of a is 1, the motor speed ranges from 0 to 1000 rpm. Note that in our model, the power amplifier, modeled by the dependent source, can deliver any current the motor requires. Of course, this is not possible, but it does demonstrate some of the tradeoffs we experience in modeling. By choosing a simple model, we were able to develop the required relationship quickly. However, other characteristics of an actual power amplifier have been omitted in this model. irwin03_101-155hr.qxd 30-06-2010 13:12 Page 139 SUMMARY 139 3.4 Design Example An 8-V source is to be used in conjunction with two standard resistors to design a voltage divider that will output 5 V when connected to a 100-A load. While keeping the consumed power as low as possible, we wish to minimize the error between the actual output and the required 5 volts. The divider can be modeled as shown in Fig. 3.32. Applying KCL at the output node yields the equation DESIGN EXAMPLE 3.23 SOLUTION VS - Vo Vo = + Io R1 R2 Using the specified parameters for the input voltage, desired output voltage, and the current source, we obtain R1 = 3R2 5 + (100)R2 By trial and error, we find that excellent values for the two standard resistors are R1 = 10 k and R2 = 27 k. Large resistor values are used to minimize power consumption. With this selection of resistors the output voltage is 5.11 V, which is a percent error of only 2.15%. Figure 3.32 A simple voltage-divider circuit with a 100-A load. R1 VS ± – + 8V R2 Vo - • Io 100 A SUMMARY A voltage source—independent or dependent—may be connected between a nonreference node and the reference node or between two nonreference nodes. A supernode is formed by a voltage source and its two connecting nonreference nodes. Nodal Analysis for an N-node Circuit ■ ■ Determine the number of nodes in the circuit. Select one node as the reference node. Assign a node voltage between each nonreference node and the reference node. All node voltages are assumed positive with respect to the reference node. For an N-node circuit, there are N 1 node voltages. As a result, N 1 linearly independent equations must be written to solve for the node voltages. Write a constraint equation for each voltage source— independent or dependent—in the circuit in terms of the assigned node voltages using KVL. Each constraint equation represents one of the necessary linearly independent equations, and Nv voltage sources yield Nv linearly independent equations. For each dependent voltage source, express the controlling variable for that source in terms of the node voltages. ■ Use KCL to formulate the remaining N - 1 - Nv linearly independent equations. First, apply KCL at each nonreference node not connected to a voltage source. Second, apply KCL at each supernode. Treat dependent current sources like independent current sources when formulating the KCL equations. For each dependent current source, express the controlling variable in terms of the node voltages. Loop Analysis for an N-loop Circuit ■ Determine the number of independent loops in the circuit. Assign a loop current to each independent loop. For an N-loop circuit, there are N-loop currents. As a result, • irwin03_101-155hr.qxd 140 ■ ■ 30-06-2010 CHAPTER 3 13:12 Page 140 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S N linearly independent equations must be written to solve for the loop currents. If current sources are present in the circuit, either of two techniques can be employed. In the first case, one loop current is selected to pass through one of the current sources. The remaining loop currents are determined by opencircuiting the current sources in the circuit and using this modified circuit to select them. In the second case, a current is assigned to each mesh in the circuit. Write a constraint equation for each current source— independent or dependent—in the circuit in terms of the ■ assigned loop currents using KCL. Each constraint equation represents one of the necessary linearly independent equations, and NI current sources yield NI linearly independent equations. For each dependent current source, express the controlling variable for that source in terms of the loop currents. Use KVL to formulate the remaining N - NI linearly independent equations. Treat dependent voltage sources like independent voltage sources when formulating the KVL equations. For each dependent voltage source, express the controlling variable in terms of the loop currents. PROBLEMS • 3.1 Find I1 in the circuit in Fig. P3.1. 3.4 Use nodal analysis to find V1 in the circuit in Fig P3.4. 2 k⍀ 4 mA 2 mA + 3 k⍀ 6 mA 4 mA 4 k⍀ V1 3 k⍀ 2 k⍀ - I1 Figure P3.1 2 k⍀ 4 mA Figure P3.4 3.5 Find V1 and V2 in the circuit in Fig. P3.5 using nodal analysis. 3.2 Find I1 in the network in Fig. P3.2. 6 mA 6 mA 6 k⍀ 12 k⍀ 4 k⍀ 4 k⍀ I1 + V1 4 mA Figure P3.2 + 4 k⍀ V2 6 k⍀ - 3 k⍀ 6 k⍀ - Figure P3.5 3.3 Find Io in the circuit in Fig. P3.3. 3.6 Use nodal analysis to find both V1 and Vo in the circuit in Fig 3.6. 6 mA 2 mA 2 k⍀ 1 k⍀ V1 V2 6 k⍀ 4 mA 3 k⍀ 2 mA 12 mA Io Figure P3.3 3 k⍀ + 2 k⍀ 6 k⍀ 1 k⍀ Vo - Figure P3.6 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 141 PROBLEMS 3.7 Find Io in the circuit in Fig. P3.7 using nodal analysis. Io 3.11 Use nodal analysis to find Io in the network in Fig. P3.11. 2 k⍀ 2 mA 4 mA 1 k⍀ 2 k⍀ 6 mA 4 mA 4 k⍀ 2 k⍀ Figure P3.7 4 k⍀ 3.8 Find Io in the network in Fig. P3.8 using nodal analysis. 6 mA 12 k⍀ Io 4 mA Figure P3.11 2 k⍀ 2 mA 1 k⍀ 6 mA 1 k⍀ Io 3.12 Find Vo in the network in Fig. P3.12 using nodal analysis. Figure P3.8 + Vo - 3.9 Find Io in the circuit in Fig. P3.9. 6 k⍀ 4 mA 2 k⍀ ± – 12 V 1 k⍀ Io 2 mA 12 k⍀ ± – 6 k⍀ 6V 6 mA Figure P3.12 2 k⍀ Figure P3.9 3.10 Find Io in the circuit in Fig. P3.10 using nodal analysis. 3.13 Find Vo in the circuit in Fig. P3.13 using nodal analysis. 8 k⍀ 1 k⍀ 2 mA 2 k⍀ 2 k⍀ 2 k⍀ 1 mA + 3 k⍀ 1 k⍀ 6 k⍀ 12 V 1 k⍀ Vo – Io Figure P3.10 + – Figure P3.13 141 irwin03_101-155hr.qxd 142 30-06-2010 CHAPTER 3 13:12 Page 142 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.14 Use nodal analysis to find Vo in the circuit in Fig. P3.14. 3.18 Find Vo in the circuit in Fig. P3.18 using nodal analysis. 4 k⍀ 2 k⍀ + + 12 V ± – 2 mA Vo 2 k⍀ 1 k⍀ 2 mA - Vo 2 k⍀ 2 k⍀ Figure P3.14 12 V + – 1 k⍀ – 3.15 Find Io in the network in Fig. P3.15 using nodal Figure P3.18 analysis. 1 k⍀ 1 k⍀ Io 2 k⍀ 3.19 Find Io in the circuit in Fig. P3.19 using nodal analysis. + 4 mA 2 mA 6V 12 V Figure P3.15 + – 2 k⍀ Io 1 k⍀ 2 k⍀ 3.16 Find Io in the circuit in Fig. P3.16 using nodal analysis. 2 k⍀ 2 k⍀ 1 k⍀ 2 mA 1 k⍀ Figure P3.19 12 V + – 2 mA + 6V Io Figure P3.16 3.20 Find Vo in the network in Fig. P3.20 using nodal analysis. 3.17 Find Vo in the circuit in Fig. P3.17 using nodal analysis. 12 V –+ 1 k⍀ + – 1 k⍀ 2 k⍀ 2 k⍀ 4 mA + + 2 k⍀ 4 mA 1 k⍀ Vo 1 k⍀ 2 k⍀ Vo – – Figure P3.17 12 V Figure P3.20 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 143 PROBLEMS 3.21 Find Vo in the circuit in Fig. P3.21 using nodal analysis. 143 3.24 Find Io in the circuit in Fig. P3.24 using nodal analysis. 1 k⍀ 2 k⍀ 1 k⍀ 4V 2 k⍀ –+ + 2 k⍀ 12 V + 4 mA Vo 1 k⍀ + - 12 V – Figure P3.21 + 1 k⍀ 6V Io Figure P3.24 3.25 Use nodal analysis to solve for the node voltages in the 3.22 Find Vo in the network in Fig. P3.22 using nodal analysis. circuit in Fig. P3.25. Also calculate the power supplied by the 2-mA current source. 12 k⍀ + – 2 k⍀ + – 4V 6 k⍀ 4 k⍀ 2 k⍀ 12 V + 4 mA Vo 2 k⍀ + - + 2 mA 12 V 24 V – Figure P3.25 Figure P3.22 3.26 Use nodal analysis to determine the node voltages defined in the circuit in Fig. P3.26. 3.23 Find Io in the circuit in Fig. P3.23 using nodal analysis. V1 0.5 S 1 k⍀ 24 V 2 k⍀ 12 V + – + – 6V Figure P3.23 1 k⍀ 5 mA V3 V2 2S 0.5 S 2 k⍀ + Io Figure P3.26 1S 3 mA V4 + – 12 V irwin03_101-155hr.qxd 144 30-06-2010 CHAPTER 3 13:12 Page 144 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.27 Use nodal analysis to solve for the node voltages in the 3.31 Find Vo in the network in Fig. P3.31. circuit in Fig. P3.27. Also calculate the power supplied by the 2-mA current source. 2 mA 4 k⍀ ± – 2 k⍀ ± – 6V 12 V 2 k⍀ 1 k⍀ + - 2 k⍀ + 2V + 2 k⍀ 2 k⍀ + – 1 k⍀ 4V 1 k⍀ 4 k⍀ + – 12 V + – 8V Vo - 16 V Figure P3.31 Figure P3.27 3.32 Find Io in the network in Fig. P3.32 using nodal analysis. 3.28 Use nodal analysis to find Vo in the network in Fig. P3.28. 1 k⍀ + 1 k⍀ 2 k⍀ 2 k⍀ 12 V 1 k⍀ + – – Vx 2 mA 2 Vx Io Figure P3.32 Vo 1 k⍀ + – 2 k⍀ 12 V + 1 k⍀ ± – 1 k⍀ - 3.33 Find Io in the network in Fig. P3.33 using nodal analysis. Figure P3.28 3.29 Use nodal analysis to find Vo in the circuit in Fig. P3.29. ± – 2 k⍀ 6V + 1 k⍀ 1 k⍀ 1 k⍀ 2 mA 12 V Io – + 4 Ix Vo - 3.34 Find Vo in the network in Fig. P3.34 using nodal analysis. Figure P3.29 1 k⍀ 3.30 Use nodal analysis to find Vo in the circuit in Fig. P3.30. + 6 k⍀ 1 k⍀ Figure P3.33 + 12 V 1 k⍀ 2 k⍀ + – 2 mA 1 k⍀ 4 mA Ix 6 k⍀ Figure P3.30 2 k⍀ + ±– Vo - 2 k⍀ 6 k⍀ 12 V 6 k⍀ + – 2 Vo 6 k⍀ 2 mA 1 k⍀ Vo – Figure P3.34 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 145 145 PROBLEMS 3.35 Find Vo in the circuit in Fig. P3.35 using nodal 3.38 Find Io in the network in Fig. P3.38 using nodal analysis. analysis. 1 k⍀ Ix 1 k⍀ + – 2 Vx 4 Ix 2 mA + Vx 1 k⍀ + + – 12 V 1 k⍀ 1 k⍀ – 1 k⍀ Vo + – 1 k⍀ 6V Io – Figure P3.38 Figure P3.35 3.39 Find Vo in the network in Fig. P3.39 using nodal analysis. + + 12 V – Ix 3.36 Find Vo in the network in Fig. P3.36 using nodal analysis. 2 mA Vo 1 k⍀ 1 k⍀ + 2 Ix 1 k⍀ 1 k⍀ 1 k⍀ – Vo 1 k⍀ Figure P3.39 1 k⍀ 12 V + – 2 Io 3.40 Use nodal analysis to find Vo in the circuit in Io Fig. P3.40. – 12 V Figure P3.36 + 4 Ix Vx +– Vo 1 k⍀ 2 mA 1 k⍀ 1 k⍀ 2 Vx + – Ix – Figure P3.40 3.37 Find Io in the network in Fig. P3.37 using nodal analysis. 3.41 Use nodal analysis to find Vo in the circuit in Fig. P3.41. 4 mA Io 1 k⍀ 1 k⍀ 4 Vx + – 2 mA 1 k⍀ + Vx + – –+ 6V 6V + – 2 Vx Vx – – Figure P3.37 + Figure P3.41 + Vo – irwin03_101-155hr3.qxd 146 2-08-2010 CHAPTER 3 16:37 Page 146 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.42 Use nodal analysis to find Vo in the network in 3.45 Find Io in the circuit in Fig. P3.45 using nodal Fig. P3.42. analysis. 4 mA 12 k⍀ 12 k⍀ 1 k⍀ 1 k⍀ –+ + – 2 Vx 2 mA 1 k⍀ Vx 1 k⍀ – Vo 12 k⍀ Ix + + 12 V 4Ix 12 mA 1 k⍀ Io Figure P3.45 – Figure P3.42 3.46 Find Vo in the circuit in Fig. P3.46. + Vx 3.43 Use nodal analysis to find Vo in the circuit in 1 k⍀ + 2 k⍀ 1 k⍀ 2Vx — 1000 2 mA Vo 1 k⍀ - - Fig. P3.43. Figure P3.46 2 Vx –+ 12 V 1 k⍀ 1 k⍀ –+ 1 k⍀ + + 4 mA Vx 2 mA 1 k⍀ 1 k⍀ – 3.47 Use nodal analysis to find Vo in the circuit in Fig. P3.47. In addition, find all branch currents and check your answers using KCL at every node. Vo – + 2000Ix 12 k⍀ Figure P3.43 6V + + 2 k⍀ 4 k⍀ 2 mA 4 k⍀ Ix Vo - Figure P3.47 3.44 Use nodal analysis to find Io in the circuit in Fig. P3.44. 3.48 Determine Vo in the network in Fig. P3.48 using nodal V1 4 Ix 1 k⍀ V2 1 k⍀ + Vx + – analysis. 1 k⍀ – + – Ix 2 Vx 2Ix 2 mA 1 k⍀ V3 Ix 1 k⍀ + 1 k⍀ 6V 1 k⍀ Io 2 mA 1 k⍀ 12 V + 1 k⍀ Vo - Figure P3.44 Figure P3.48 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 147 147 PROBLEMS 3.49 Use nodal analysis to find V1, V2, V3, and V4 in the 3.52 Use nodal analysis to determine the node voltages defined circuit in Fig. P3.49. in the circuit in Fig. P3.52. 12 V 2Vx ±– V2 V1 6⍀ 10 ⍀ ± 3Ix V1 1⍀ ± – Ix 6⍀ V3 8⍀ V4 4⍀ 3A – VA –+ 4A 2 VA Vx 5⍀ 3A 8⍀ V2 – 3⍀ 12 V V3 + – 4⍀ V4 4A + Figure P3.49 Figure P3.52 3.50 Use nodal analysis to determine the node voltages defined 3.53 Find Io in the network in Fig. P3.53 using mesh analysis. in the circuit in Fig. P3.50. V4 IA – VA 6 k⍀ 3⍀ + – 12 ⍀ + 2 VA V2 V1 + – Io V5 6⍀ 9⍀ 5A V3 + ± – 6 k⍀ 24 V Figure P3.53 4 IA 7⍀ 12 V 6V 4 k⍀ 3.54 Find Io in the circuit in Fig. P3.54. 4 k⍀ 4 k⍀ Figure P3.50 2 k⍀ 24 V ± – Io 3.51 Use nodal analysis to determine the node voltages defined ± – in the circuit in Fig. P3.51. 2 k⍀ 6V V5 Ix – Figure P3.54 4⍀ 5⍀ V1 V4 Vx – + 0.5 Vx + 2 Ix 3A 8⍀ V2 6 ⍀ 3 k⍀ 3.55 Find Vo in the network in Fig. P3.55 using mesh analysis. 4V V3 + – 12 V ±- + + 12 V 4 k⍀ 4 k⍀ 4 k⍀ Vo - Figure P3.51 Figure P3.55 irwin03_101-155hr.qxd 148 30-06-2010 CHAPTER 3 13:12 Page 148 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.56 Find Io in the circuit in Fig. P3.56. 6 k⍀ Fig. P3.60. 12 k⍀ -± ± – 3V 3.60 Use mesh analysis to find Vo in the network in 6V 12 k⍀ – ± 6 k⍀ 3V 12 V + 2 k⍀ 3 k⍀ ± – 6 k⍀ 2 mA Vo - Io Figure P3.60 Figure P3.56 3.61 Find Vo in the network in Fig. P3.61 using loop analysis. 3.57 Use mesh analysis to find the power dissipated in the 6-k resistor in Fig. P3.57. 2 k⍀ 4 k⍀ 6V + – 8 k⍀ 2 k⍀ + – 2 k⍀ 4 mA + 1 k⍀ 8V – 2 k⍀ 4 k⍀ 2 k⍀ + 12 V – Vo + – 6 k⍀ 1 k⍀ 6 mA 16 V Figure P3.61 Figure P3.57 3.62 Find Vo in the circuit in Fig. P3.62 using loop analysis. 2 mA 3.58 Find Vo in the circuit in Fig. P3.58 using mesh analysis. -± 4 mA 6V Vo 1 k⍀ + 1 k⍀ 12 V –+ – 1 k⍀ 1 k⍀ 6V ± – 2 k⍀ 2 k⍀ 1 k⍀ + Vo Figure P3.62 - 3.63 Find Io in the circuit in Fig. P3.63 using loop analysis. Figure P3.58 2 mA 3.59 Use loop analysis to find Vo in the network in Fig. P3.59. 2 k⍀ 1 k⍀ 2 k⍀ 2 mA 2 k⍀ –± 12 V 1 k⍀ + Vo 4 mA 2 k⍀ Io - Figure P3.59 1 k⍀ Figure P3.63 – + 6V irwin03_101-155hr.qxd 30-06-2010 13:12 Page 149 PROBLEMS 3.64 Find Io in the network in Fig. P3.64 using loop analysis. 3.68 Find Io in the network in Fig. P3.68 using loop analysis. 2 mA 6 k⍀ 2 k⍀ 6 k⍀ 1 k⍀ 1 k⍀ 149 4 mA 6V 2 k⍀ 6 k⍀ + 5 mA 6 k⍀ Io Figure P3.68 Io Figure P3.64 3.65 Find Vo in the network in Fig. P3.65 using loop analysis. 3.69 Use loop analysis to find Vo in the circuit in Fig. P3.69. + 2 k⍀ 6 mA 1 k⍀ + 2 mA 6V –+ 12 k⍀ +– Vo 4V 4 mA 1 k⍀ 12 V + Vo - 6 k⍀ 12 k⍀ 2 mA 2 k⍀ – Figure P3.65 Figure P3.69 3.70 Using loop analysis, find Vo in the network in 3.66 Find Vo in the circuit in Fig. P3.66 using loop analysis. Fig. P3.70. + – + 6 mA 2 k⍀ 2 k⍀ 4V 1 k⍀ 4 mA ±– 2 k⍀ Vo 1 k⍀ –+ 6V 1 k⍀ 6V + + 12 V 1 k⍀ 2 k⍀ Vo 1 k⍀ 4 mA - – Figure P3.70 Figure P3.66 3.67 Find Io in the network in Fig. P3.67 using loop analysis. 3.71 Find Io in the circuit in Fig. P3.71. Io 1 k⍀ 6V 12 V 1 k⍀ 1 k⍀ 1 k⍀ ±– 6 mA + - 2 k⍀ 2 mA 6 k⍀ Io Figure P3.67 Figure P3.71 4 k⍀ 2 k⍀ 1 mA 2 k⍀ irwin03_101-155hr.qxd 150 30-06-2010 CHAPTER 3 13:12 Page 150 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.72 Use loop analysis to find Io in the network in Fig. P3.72. 3.76 Using loop analysis, find Io in the circuit in Fig. P3.76. 1 k⍀ 4 mA 1 k⍀ + - 12 V 1 k⍀ 2 mA 12 V 1 k⍀ Io 1 k⍀ 1 k⍀ 2 mA 1 k⍀ 1 k⍀ 4 mA 1 k⍀ + 1 k⍀ 1 k⍀ 1 k⍀ 1 k⍀ Io 1 k⍀ Figure P3.76 Figure P3.72 3.73 Find Io in the circuit in Fig. P3.73 using loop analysis. 3.77 Find the mesh currents in the network in Fig. P3.77. 4 mA + – 1 k⍀ 1 k⍀ 2 k⍀ 6V 1 k⍀ 2 k⍀ 6 mA Io 2 mA 1 k⍀ 2 mA 1 k⍀ 6V + - Figure P3.73 3.74 Find Io in the network in Fig. P3.74 using loop analysis. 2 mA 1 k⍀ I1 I2 2 k⍀ 2 k⍀ + 12 V I5 1 k⍀ 1 k⍀ 1 k⍀ I3 I4 Figure P3.77 1 k⍀ 3.78 Use loop analysis to find Vo in the network in Fig. P3.78. +– –+ 6V 6 mA Ix 4 k⍀ + 12 V 4 mA 2 k⍀ + 12 V 1 k⍀ 2 k⍀ Io 2 k⍀ 2 Ix - Figure P3.74 3.75 Find Vo in the circuit in Fig. P3.75 using loop analysis. 4 mA 2 k⍀ 1 k⍀ + Vo 1 k⍀ 2 mA Figure P3.75 + – 6V – + 4V 2 mA Figure P3.78 3.79 Find Vo in the circuit in Fig. P3.73 using nodal analysis. – 1 k⍀ 1 k⍀ 4 mA 2 k⍀ Vo Vo — 2 ± – ±- + 12 V 1 k⍀ 2 k⍀ Vo - Figure P3.79 irwin03_101-155hr.qxd 30-06-2010 13:12 Page 151 151 PROBLEMS 3.84 Find Vo in the network in Fig. P3.84 using nodal 3.80 Use nodal analysis to find Vo in Fig. P3.80. analysis. Ix + 2Ix 3 mA 10 k⍀ 10 k⍀ Vo - Ix Figure P3.80 + 10 k⍀ 4000Ix + 10 k⍀ 4 mA 10 k⍀ Vo - Figure P3.84 3.85 Find Io in the circuit in Fig. P3.85 using loop analysis. 2 Ix 3.81 Find the power supplied by the 2-A current source in the network in Fig. P3.81 using loop analysis. Io Ix 4⍀ + - 6V 4⍀ 10 V 5⍀ 2 Ix 2A 1 k⍀ 1 k⍀ + – 2 k⍀ 2 mA Ix Figure P3.85 Figure P3.81 3.86. Use mesh analysis to find Vo in the circuit in Fig. P3.86. 6 Vx ±– 3.82 Find Io in the network in Fig. P3.82. 6 mA 8 k⍀ ±– 4000Ix 2 k⍀ 2 k⍀ + 4 k⍀ 12 k⍀ Ix Vx 12 k⍀ 12 mA 4 k⍀ Io + 12 k⍀ - Vo - Figure P3.86 Figure P3.82 3.87 Using mesh analysis, find Vo in the circuit in Fig. P3.87. + 3.83 Find Vo in the circuit in Fig. P3.83 using loop analysis. Vx 2 k⍀ — 4000 + 12 V Ix 1 k⍀ 1 k⍀ 4 k⍀ + 1 k⍀ 1 k⍀ 4 k⍀ + Vo 6 mA Ix Figure P3.83 - Vx - Figure P3.87 Vo 4 k⍀ - irwin03_101-155hr.qxd 152 30-06-2010 CHAPTER 3 13:12 Page 152 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.88 Find Vo in the network in Fig. P3.88 . Vx - 3.92 Using loop analysis, find Vo in the circuit in Fig. P3.92. + + 1 k⍀ 12 mA 1 k⍀ 1 k⍀ Vo 1 k⍀ 1 k⍀ 12 V + - + 1 k⍀ Vx 2 k⍀ 2 k⍀ 4 mA 4 mA + + 2 k⍀ 1 k⍀ Vx — 1000 1 k⍀ 1 k⍀ 1 k⍀ 2Vx 1 k⍀ - Figure P3.88 Vo - Figure P3.92 3.89 Using loop analysis, find Vo in the circuit in Fig. P3.89. + 2 k⍀ 1 k⍀ 1 k⍀ 2 k⍀ 12 V 3.93 Using loop analysis, find Io in the network in Fig. P3.93. + 2 mA 1 k⍀ 2Vx + + - Vx Vo - 1 k⍀ + 1 k⍀ Figure P3.89 Ix 1 k⍀ Io 12 V 1 k⍀ + Vx - 2 mA 1 k⍀ Ix 1 k⍀ Figure P3.93 + 1 k⍀ 12 V ± – 2Ix 2 mA 1 k⍀ 1 k⍀ 1 k⍀ 1 k⍀ 3.90 Using loop analysis, find Vo in the circuit in Fig. P3.90. 2Ix 2Vx + Vo 1 k⍀ - Figure P3.90 3.94 Use loop analysis to find Io in the circuit in Fig. P3.94. 3.91 Using loop analysis, find Vo in the network in Fig. P3.91. 2 Ix 4 mA 1 k⍀ 1 k⍀ + 12 V 1 k⍀ 2Vo ± – + 1 k⍀ Vo - Figure P3.91 – 1 k⍀ 1 k⍀ + 2 k⍀ Vx 6V + – Io Figure P3.94 1 k⍀ Ix + – 2 Vx irwin03_101-155hr.qxd 30-06-2010 13:12 Page 153 153 PROBLEMS 3.95 Find Ix in the circuit in Fig. P3.95 using loop analysis. 3.98 Solve for the mesh currents defined in the circuit in Fig. P3.98. 2⍀ 2 mA 4⍀ IA + – Vx I3 1⍀ 1 k⍀ Ix 2 k⍀ Ix + – I4 8⍀ + Vx – 1 k⍀ 2VA – 5⍀ + VA + – 9V 2A 3IA I1 2⍀ I2 Figure P3.95 Figure P3.98 3.99 Solve for the mesh currents defined in the circuit in Fig. P3.99. 2⍀ 3.96 Find Io in the circuit in Fig. P3.96 using loop analysis. 4⍀ + Ix 2 Vx 4⍀ + – + – 1 k⍀ + Vx 6V – Vx I1 5⍀ – 3Ix 1 k⍀ 1 k⍀ 2 mA Io 2Vx – + I3 Figure P3.96 3⍀ + – 7⍀ 2 Ix 2 k⍀ I2 8⍀ 2⍀ + – I4 12 V Ix Figure P3.99 3.100 Solve for the mesh currents defined in the circuit in Fig. P3.100. 3.97 Find Vo in the circuit in Fig. P3.97 using loop analysis. 10 V 2⍀ +– 2 k⍀ 2 Vx +– 2 Ix + 1 k⍀ Vo – Figure P3.97 + – Vx + 2 mA 1 k⍀ I4 Vx – 7⍀ 2⍀ 4⍀ 1 k⍀ 2 k⍀ Ix 4 mA 20 V + – I1 Figure P3.100 I2 5A I3 2Vx 8⍀ irwin03_101-155hr.qxd 154 30-06-2010 CHAPTER 3 13:12 Page 154 N O D A L A N D L O O P A N A LY S I S T E C H N I Q U E S 3.101 Using loop analysis, find Io in the circuit in Fig. P3.101. 3.104 Use both nodal and loop analyses to determine Io in the circuit in Fig. P3.104. ± – 1 k⍀ 12 V 1 k⍀ 1 k⍀ Io 2Vx – ± 6 mA Ix 1 k⍀ 1 k⍀ + 1 k⍀ 1 k⍀ + 2Ix 1 k⍀ - Ix 1 k⍀ 1 k⍀ 1 k⍀ Io 1 k⍀ Vx 1 k⍀ 2Vx ± – 2 Ix 2 mA 12 V + Vx - 1 k⍀ Figure P3.101 Figure P3.104 3.102 Use mesh analysis to determine the power delivered by the indepenent 3-V source in the network in Fig. P3.102. 100 ⍀ 200 ⍀ 3.105 Use both nodal and loop analyses to find Vo in the circuit in Fig. P3.105. 40 mA 6 Vx ± – 1 k⍀ 300 ⍀ 3V 100 ⍀ ± – 600 ⍀ - + 12 V + - + + + 2 k⍀ 2 k⍀ 2 k⍀ 4 mA Vx 1 k⍀ 1 k⍀ 1 k⍀ 2Vx Vx Vo 1 k⍀ - Figure P3.102 - Figure P3.105 3.103 Use mesh analysis to find the power delivered by the 3.106 Find Io in the network in Fig. P3.106 using nodal current-control voltage source in the circuit in Fig. P3.103. analysis. 6⍀ Vx + 3A 6⍀ ± – 1 k⍀ 15 Ix — 8 Ix 32 ⍀ 1⍀ 2Vx – ± Figure P3.106 6 mA 1 k⍀ 1 k⍀ 1 k⍀ + 2Ix 1 k⍀ Ix Figure P3.103 1 k⍀ Io + Vx 12 V 1 k⍀ Vx - irwin03_101-155hr.qxd 30-06-2010 13:12 Page 155 TYPICAL PROBLEMS FOUND ON THE FE EXAM 155 TYPICAL PROBLEMS FOUND ON THE FE EXAM • 3FE-4 Determine the voltage Vo in the circuit in 3FE-1 Find Vo in the circuit in Fig. 3PFE-1. Fig. 3PFE-4. a. 3.33 V c. 9.33 V a. 3.28 V c. 6.43 V b. 8.25 V d. 2.25 V b. 4.14 V d. 2.25 V Vx 2⍀ 12 V 1⍀ 12 V ± – + 4⍀ + + 4⍀ Vx + 6⍀ 2⍀ 4⍀ Ix Vo 2Ix - 6V Figure 3PFE-4 2⍀ Vo - 3FE-5 What is the voltage V1 in the circuit in Fig. 3PFE-5? Figure 3PFE-1 a. 7 V c. 2 V b. 5 V d. 4 V 3FE-2 Determine the power dissipated in the 6-ohm resistor in V1 the network in Fig. 3PFE-2. a. 8.2 W c. 4.4 W b. 15.3 W d. 13.5 W 4⍀ 1⍀ 3⍀ 8A Vx I1 12 V ± – 12 ⍀ 6⍀ 2 I1 Figure 3PFE-2 3FE-3 Find the current Ix in the 4-ohm resistor in the circuit in Fig. 3PFE-3. a. 20 A c. 7 A b. 12 A d. 14 A 12 V –± 6⍀ 2A 3⍀ + Vx - Figure 3PFE-3 4⍀ Ix ± – 2Vx Figure 3PFE-5 –± 10 V 4A 2⍀ ± – 15 V irwin04-156-188hr.qxd 9-07-2010 14:17 Page 156 4 C H A PT E R OPERATIONAL AMPLIFIERS THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Learn how to model the op-amp device ■ Learn how to analyze a variety of circuits that employ op-amps ■ Understand the use of op-amps in a number of practical applications Courtesy of NewsCom R Robotic Surgery Robotic-assisted surgery may well be the practice sessions beforehand on computer simulations, most significant story in medical circles since the Jarvik artifi- further increasing their accuracy in a field where precision is cial heart hit the headlines years ago. Using small surgical literally a matter of life or death. cuts, this minimally invasive procedure leads to faster recov- In this chapter, you will study the operational amplifier, ery and less pain, bleeding, and risk of infection. You can or op-amp, a key building block in robotics. Nodal analysis expect to leave the hospital after an overnight stay and walk and Ohm’s law are all that you need to analyze this basic within a day following kidney or bladder removal, once con- electronic component. An active device that receives external sidered major surgeries. Robotic-assisted techniques have power, the op-amp appears in common configurations for even been extended to certain heart surgeries. inverting, amplifying, and summing voltage signals. The Robotics allows surgeons to work at a console only a few op-amp also provides design separation for cascaded stages. feet away from the operating table, remotely manipulating Surgeons at the forefront of healthcare rely on electronic robotic arms at the bedside. The primary advantage is designers for pioneering tools. With the op-amp at your reducing the impact of hand tremors on surgical instruments. command, you can begin to create designs that could prove Combining computers and robotics also lets surgeons run to be life-savers. 156 irwin04-156-188hr.qxd 9-07-2010 14:17 Page 157 SECTION 4.2 It can be argued that the operational amplifier, or op-amp as it is commonly known, is the single most important integrated circuit for analog circuit design. It is a versatile interconnection of transistors and resistors that vastly expands our capabilities in circuit design, from engine control systems to cellular phones. Early op-amps were built of vacuum tubes, making them bulky and power hungry. The invention of the transistor at Bell Labs in 1947 allowed engineers to create op-amps that were much smaller and more efficient. Still, the op-amp itself consisted of individual transistors and resistors interconnected on a printed circuit board (PCB). When the manufacturing process for integrated circuits (ICs) was developed around 1970, engineers could finally put all of the op-amp transistors and resistors onto a single IC chip. Today, it is common to find as many as four high-quality op-amps on a single IC for as little as $0.40. A sample of commercial op-amps is shown in Fig. 4.1. Why are they called operational amplifiers? Originally, the op-amp was designed to perform mathematical operations such as addition, subtraction, differentiation, and integration. By adding simple networks to the op-amp, we can create these “building blocks” as well as voltage scaling, current-to-voltage conversion, and myriad more complex applications. (a) OP-AMP MODELS 157 4.1 Introduction (b) Figure 4.1 A selection of op-amps. On the left (a) is a discrete op-amp assembled on a printed circuit board (PCB). On the right, top-down, a LM324 DIP, LMC6492 DIP, and MAX4240 in a SO-5 package (small outline/5 pins). The APEX PA03 with its lid removed (b) showing individual transistors and resistors. (Left, Courtesy of Mark Nelms and Jo Ann Loden; right, Courtesy of Milt Perrin, Apex Microtechnology Corp.) How can we, understanding only sources and resistors, hope to comprehend the performance of the op-amp? The answer lies in modeling. When the bells and whistles are removed, an op-amp is just a really good voltage amplifier. In other words, the output voltage is a scaled replica of the input voltage. Modern op-amps are such good amplifiers that it is easy to create an accurate, first-order model. As mentioned earlier, the op-amp is very popular and is used extensively in circuit design at all levels. We should not be surprised to find that op-amps are available for every application—low voltage, high voltage, micropower, high speed, high current, and so forth. Fortunately, the topology of our model is independent of these issues. We start with the general-purpose LM324 quad (four in a pack) op-amp from National Semiconductor, shown in the upper right corner of Fig. 4.1a. The pinout for the LM324 is shown in Fig. 4.2 for a DIP (Dual Inline Pack) style package with dimensions in inches. Recognizing there are four identical op-amps in the package, we will focus on amplifier 1. Pins 3 and 2 are the input pins, IN 1+ and IN 1-, and are called the noninverting and inverting inputs, respectively. The output is at pin 1. A relationship exists between the output and input voltages, Vo = A o(IN+ - IN-) 4.1 4.2 Op-Amp Models irwin04-156-188hr.qxd 158 9-07-2010 CHAPTER 4 14:17 Page 158 O P E R AT I O N A L A M P L I F I E R S Figure 4.2 The pinout (a) and dimensional diagram (b) of the LM324 quad op-amp. Note the pin pitch (distance pin-to-pin) is 0.1 inches—standard for DIP packages. 0.78 14 13 12 11 10 9 8 OUT 4 IN 4– IN 4+ VEE IN 3– IN 3+ IN 3 0.04 4 – ± –3 ± 0.3 1 ± – ±2 – OUT 1 IN 1– IN 1+ 1 2 3 0.1 0.06 VCC IN 2– IN 2+ OUT 2 4 5 6 7 (a) (b) where all voltages are measured with respect to ground and A o is the gain of the op-amp. (The location of the ground terminal will be discussed shortly.) From Eq. (4.1), we see that when IN+ increases, so will Vo . However, if IN- increases, then Vo will decrease—hence the names noninverting and inverting inputs. We mentioned earlier that op-amps are very good voltage amplifiers. How good? Typical values for A o are between 10,000 and 1,000,000! Amplification requires power that is provided by the dc voltage sources connected to pins 4 and 11, called VCC and VEE , respectively. Fig. 4.3 shows how the power supplies, or rails, are connected for both dual- and single-supply applications and defines the ground node to which all input and output voltages are referenced. Traditionally, VCC is a positive dc voltage with respect to ground, and VEE is either a negative voltage or ground itself. Actual values for these power supplies can vary widely depending on the application, from as little as one volt up to several hundred. How can we model the op-amp? A dependent voltage source can produce Vo ! What about the currents into and out of the op-amp terminals (pins 3, 2, and 1)? Fortunately for us, the currents are fairly proportional to the pin voltages. That sounds like Ohm’s law. So, we model the I-V performance with two resistors, one at the input terminals (Ri) and another at the output (Ro). The circuit in Fig. 4.4 brings everything together. Figure 4.3 Schematics showing the power supply connections and ground location for (a) dual-supply and (b) single-supply implementations. Figure 4.4 A simple model for the gain characteristics of an op-amp. IN+ IN– ± – VCC VCC IN+ out IN– VEE VEE ± – VCC VCC VEE (a) (b) + + vin(t) IN+(t) + - IN–(t) - Ro + Ri Aovin ± – vo(t) - irwin04-156-188hr.qxd 9-07-2010 14:17 Page 159 SECTION 4.2 RTh1 + vin(t) Ri - ± Aovin – RL vo(t) - What values can we expect for A o , Ri , and Ro ? We can reason through this issue with the help of Fig. 4.5 where we have drawn an equivalent for the circuitry that drives the input nodes and we have modeled the circuitry connected to the output with a single resistor, RL . Since the op-amp is supposed to be a great voltage amplifier, let’s write an equation for the overall gain of the circuit Vo兾VS . Using voltage division at the input and again at the output, we quickly produce the expression Vo Ri RL = c dA c d VS Ri + RTh1 o Ro + RL To maximize the gain regardless of the values of RTh1 and RL , we make the voltage division ratios as close to unity as possible. The ideal scenario requires that A o be infinity, Ri be infinity, and Ro be zero, yielding a large overall gain of A o . Table 4.1 shows the actual values of A o , Ri , and Ro for a sampling of commercial op-amps intended for very different applications. While A o , Ri , and Ro are not ideal, they do have the correct tendencies. The power supplies affect performance in two ways. First, each op-amp has minimum and maximum supply ranges over which the op-amp is guaranteed to function. Second, for proper operation, the input and output voltages are limited to no more than the supply voltages.* If the inputs/output can reach within a few dozen millivolts of the supplies, then the inputs/output are called rail-to-rail. Otherwise, the inputs/output voltage limits are more severe—usually a volt or so away from the supply values. Combining the model in Fig. 4.4, the values in Table 4.1, and these I/O limitations, we can produce the graph in Fig. 4.6 showing the output–input relation for each op-amp in Table 4.1. From the graph we see that LMC6492 and MAX4240 have rail-to-rail outputs while the LM324 and PA03 do not. TABLE 4.1 † A list of commercial op-amps and their model values MANUFACTURER PART NO. Ao (V/V) Ri ( M⍀ ) Ro ( ⍀ ) National LM324 100,000 1.0 20 National LMC6492 50,000 107 150 Maxim MAX4240 20,000 45 160 Apex PA03 125,000 105 2 159 Figure 4.5 Ro + VS ± – OP-AMP MODELS COMMENTS General purpose, up to 16 V supplies, very inexpensive Low voltage, rail-to-rail inputs and outputs† Micro-power (1.8 V supply @ 10 A), rail-to-rail inputs and outputs High-voltage, 75 V and highoutput current capability, 30 A. That’s 2 kW! Rail-to-rail is a trademark of Motorola Corporation. This feature is discussed further in the following paragraphs. *Op-amps are available that have input and/or output voltage ranges beyond the supply rails. However, these devices constitute a very small percentage of the op-amp market and will not be discussed here. A network that depicts an op-amp circuit. VS and RTh1 model the driving circuit, while the load is modeled by RL. The circuit in Fig. 4.4 is the op-amp model. 160 9-07-2010 CHAPTER 4 14:17 O P E R AT I O N A L A M P L I F I E R S Figure 4.6 80 15 60 10 Output voltage, vo (V) Transfer plots for the op-amps listed in Table 4.1. The supply voltages are listed in the plot legends. Note that the LMC6492 and MAX4240 have rail-torail output voltages (output voltage range extends to power supply values), while the LM324 and PA03 do not. Page 160 Output voltage, vo (V) irwin04-156-188hr.qxd 5 0 –5 40 20 0 –20 –10 LMC6492@+/– 5 V –15 LM324@+/– 15 V –40 LM324@+/– 15 V MAX4240@+/– 1.5 V PA03@+/– 75 V –60 LMC6492@+/– 5 V MAX4240@+/– 1.5 V –20 –250 –200 –150 –100 –50 0 50 –80 –1.0 –0.8 –0.6 –0.4 –0.2 0.0 100 150 200 250 Input voltage, vin (V) 0.2 0.4 0.6 0.8 1.0 Input voltage, vin (mV) Even though the op-amp can function within the minimum and maximum supply voltages, because of the circuit configuration, an increase in the input voltage may not yield a corresponding increase in the output voltage. In this case, the op-amp is said to be in saturation. The following example addresses this issue. EXAMPLE 4.1 The input and output signals for an op-amp circuit are shown in Fig. 4.7. We wish to determine (a) if the op-amp circuit is linear and (b) the circuit’s gain. SOLUTION a. We know that if the circuit is linear, the output must be linearly related, that is, proportional, to the input. An examination of the input and output waveforms in Fig. 4.7 clearly indicates that in the region t = 1.25 to 2.5 and 4 to 6 ms the output is constant while the input is changing. In this case, the op-amp circuit is in saturation and therefore not linear. b. In the region where the output is proportional to the input, that is, t = 0 to 1 ms, the input changes by 1 V and the output changes by 3.3 V. Therefore, the circuit’s gain is 3.3. Figure 4.7 An op-amp input–output characteristic. Voltage (V) Output 3 2 Input 1 0 1 2 –1 –2 –3 –4 3 4 5 6 Input Output 7 t (ms) irwin04-156-188hr.qxd 9-07-2010 14:17 Page 161 SECTION 4.2 OP-AMP MODELS Figure 4.8 + Vin VCC VS ± – Ri - ± – VS ± – + Vo - VEE Ro ± – AoVin Circuit (a) and model (b) for the unity gain buffer. + Vo - I (a) 161 (b) To introduce the performance of the op-amp in a practical circuit, consider the network in Fig. 4.8a called a unity gain buffer. Notice that the op-amp schematic symbol includes the power supplies. Substituting the model in Fig. 4.4 yields the circuit in Fig. 4.8b, containing just resistors and dependent sources, which we can easily analyze. Writing loop equations, we have VS = IRi + IRo + A o Vin Vout = IRo + A o Vin Vin = IRi Solving for the gain, Vo兾VS , we find Vo = VS 1 1 + Ri Ro + A o Ri For Ro V Ri , we have Vo L VS 1 + i– ±∞ – Figure 4.9 The origin of the name unity gain buffer should be apparent. Table 4.2 shows the actual gain values for VS = 1 V using the op-amps listed in Table 4.1. Notice how close the gain is to unity and how small the input voltage and current are. These results lead us to simplify the op-amp in Fig. 4.4 significantly. We introduce the ideal op-amp model, where A o and Ri are infinite and Ro is zero. This produces two important results for analyzing op-amp circuitry, listed in Table 4.3. TABLE 4.2 Ideal model for an operational amplifier. Model parameters: i + = i- = 0, v + = v- . TABLE 4.3 Unity gain buffer performance for the op-amps listed in Table 4.1 V in (mV) v– 1 Ao Vo L 1 VS BUFFER GAIN i+ 1 And, if A o is indeed W 1, OP-AMP v+ Consequences of the ideal op-amp model on input terminal I兾V values I ( pA ) MODEL ASSUMPTION TERMINAL RESULT LM324 0.999990 9.9999 9.9998 Ao S q input voltage S 0 V LMC6492 0.999980 19.999 1.9999 106 Ri S q input current S 0 A MAX4240 0.999950 49.998 1.1111 PA03 0.999992 7.9999 7.9999 105 From Table 4.3 we find that the ideal model for the op-amp is reduced to that shown in Fig. 4.9. The important characteristics of the model are as follows: (1) since Ri is extremely large, the input currents to the op-amp are approximately zero (i.e., i+ L i- L 0); and (2) if the output voltage is to remain bounded, then as the gain becomes very large and approaches infinity, the voltage across the input terminals must simultaneously become infinitesimally small so that as A o S q, v+ - v- S 0 (i.e., v+ - v- = 0 or v+ = v-). The difference between these input voltages is often called the error signal for the op-amp (i.e., v+ - v- = ve). irwin04-156-188hr.qxd 162 9-07-2010 CHAPTER 4 0A+ + 0V - Vin 0 A- VS ± – ± – 14:17 Page 162 O P E R AT I O N A L A M P L I F I E R S + The ground terminal shown on the op-amp is necessary for signal current return, and it guarantees that Kirchhoff’s current law is satisfied at both the op-amp and the ground node in the circuit. In summary, then, our ideal model for the op-amp is simply stated by the following conditions: Vo + i+ = i- = 0 VS - - Figure 4.10 An ideal op-amp configured as a unity gain buffer. RS + 0V - VS ± – 4.2 v+ = v- VCC 0A 0A ± – VEE + VS - Figure 4.11 A unity gain buffer with a load resistor. These simple conditions are extremely important because they form the basis of our analysis of op-amp circuits. Let’s use the ideal model to reexamine the unity gain buffer, drawn again in Fig. 4.10, where the input voltage and currents are shown as zero. Given that Vin is zero, the voltage at both op-amp inputs is VS . Since the inverting input is physically connected to the output, Vo is also VS —unity gain! Armed with the ideal op-amp model, let’s change the circuit in Fig. 4.10 slightly as shown in Fig. 4.11 where VS and RS are an equivalent for the circuit driving the buffer and RL models the circuitry connected to the output. Io There are three main points here. First, the gain is still unity. Second, the opamp requires no current from the driving circuit. Third, the output current + (Io = Vo兾RL)comes from the power supplies, through the op-amp and out of the output pin. In other words, the load current comes from the power supRL plies, which have plenty of current output capacity, rather than the driving cirVo 1 k⍀ cuit, which may have very little. This isolation of current is called buffering. An obvious question at this point is this: if Vo = VS , why not just connect VS to Vo via two parallel connection wires; why do we need to place an opamp between them? The answer to this question is fundamental and provides us with some insight that will aid us in circuit analysis and design. Consider the circuit shown in Fig. 4.12a. In this case Vo is not equal to VS because of the voltage drop across RS: Vo = VS - IRS However, in Fig. 4.12b, the input current to the op-amp is zero and, therefore, VS appears at the op-amp input. Since the gain of the op-amp configuration is 1, Vo = VS . In Fig. 4.12a the resistive network’s interaction with the source caused the voltage Vo to be less than VS . In other words, the resistive network loads the source voltage. However, in Fig. 4.12b the op-amp isolates the source from the resistive network; therefore, the voltage follower is referred to as a buffer amplifier because it can be used to isolate one circuit from another. The energy supplied to the resistive network in the first case must come from the source VS , whereas in the second case it comes from the power supplies that supply the amplifier, and little or no energy is drawn from VS . RS RS I I + VS ± – Vo Resistive network ± – VS ± – Vo - (a) Figure 4.12 Illustration of the isolation capability of a voltage follower. + - (b) Resistive network irwin04-156-188hr.qxd 9-07-2010 14:17 Page 163 SECTION 4.3 163 F U N D A M E N TA L O P - A M P C I R C U I T S As a general rule, when analyzing op-amp circuits we write nodal equations at the op-amp input terminals, using the ideal op-amp model conditions. Thus, the technique is straightforward and simple to implement. 4.3 Fundamental Op-Amp Circuits EXAMPLE Let us determine the gain of the basic inverting op-amp configuration shown in Fig. 4.13a using both the nonideal and ideal op-amp models. R1 vS R2 b – ±A a ± – 4.2 d v+ + c v– vo i+ a i– b d ±A – c - (a) (b) b a + - d d + Ro ve v+ R2 - ± – Ri b vo v– c - v– Ri vS ± – A(v+-v–) =Ave + - R1 + v+ Ro + ± – A(v -v ) + – vo c (c) (d) v1 R1 vS ± – vo - Ri R2 ve + + Ro Ave ± – vo - Figure 4.13 Op-amp circuit. (e) Our model for the op-amp is shown generically in Fig. 4.13b and specifically in terms of the parameters Ri , A, and Ro in Fig. 4.13c. If the model is inserted in the network in Fig. 4.13a, we obtain the circuit shown in Fig. 4.13d, which can be redrawn as shown in Fig. 4.13e. SOLUTION - irwin04-156-188hr.qxd 164 9-07-2010 CHAPTER 4 14:17 Page 164 O P E R AT I O N A L A M P L I F I E R S The node equations for the network are v1 v1 - vo v1 - vS + + = 0 R1 Ri R2 vo - v1 vo - Ave + = 0 R2 Ro where ve = -v1 . The equations can be written in matrix form as -a 1 vS b R2 v1 R T B R = C 1S 1 1 vo 0 + R2 Ro 1 1 1 + + R1 Ri R2 D 1 A -a b R2 Ro Solving for the node voltages, we obtain 1 1 + v1 1 R2 Ro B R = D vo 1 A ¢ R2 Ro 1 vS R2 T C R1 S 1 1 1 0 + + R1 Ri Ro where ¢ = a 1 1 1 1 1 1 1 A + + ba + b - a ba b R1 Ri R2 R2 Ro R2 R2 Ro Hence, vo = a a vS 1 A ba b R2 Ro R1 1 1 1 1 1 1 1 A + + ba + b - a ba b R1 Ri R2 R2 Ro R2 R2 Ro which can be written as vo = vS 1 - ca - AR2兾R1 B 1 1 1 1 1 1 1 A + + ba + bna b a bd R1 Ri R2 R2 Ro R2 R2 Ro If we now employ typical values for the circuit parameters (e.g., A = 105, Ri = 108 , Ro = 10 , R1 = 1 k, and R2 = 5 k), the voltage gain of the network is vo = -4.9996994 L -5.000 vS However, the ideal op-amp has infinite gain. Therefore, if we take the limit of the gain equation as A S q, we obtain vo R2 = -5.000 lim a b = S q vS R1 A Note that the ideal op-amp yielded a result accurate to within four significant digits of that obtained from an exact solution of a typical op-amp model. These results are easily repeated for the vast array of useful op-amp circuits. We now analyze the network in Fig. 4.13a using the ideal op-amp model. In this model, i+ = i- = 0 v+ = v- irwin04-156-188hr.qxd 9-07-2010 14:17 Page 165 SECTION 4.3 F U N D A M E N TA L O P - A M P C I R C U I T S 165 As shown in Fig. 4.13a, v+ = 0 and, therefore, v- = 0. If we now write a node equation at the negative terminal of the op-amp, we obtain vS - 0 vo - 0 + = 0 R1 R2 or vo R2 = vS R1 and we have immediately obtained the results derived previously. Notice that the gain is a simple resistor ratio. This fact makes the amplifier very versatile in that we can control the gain accurately and alter its value by changing only one resistor. Also, the gain is essentially independent of op-amp parameters. Since the precise values of A o , Ri , and Ro are sensitive to such factors as temperature, radiation, and age, their elimination results in a gain that is stable regardless of the immediate environment. Since it is much easier to employ the ideal op-amp model rather than the nonideal model, unless otherwise stated we will use the ideal op-amp assumptions to analyze circuits that contain operational amplifiers. Problem-Solving Strategy Step 1. Use the ideal op-amp model: A o = q, Ri = q, Ro = 0. ● i+ = i- = 0 Op-Amp Circuits v+ = v- ● Step 2. Apply nodal analysis to the resulting circuit. Step 3. Solve nodal equations to express the output voltage in terms of the op-amp input signals. Let us now determine the gain of the basic noninverting op-amp configuration shown in Fig. 4.14. vin ± – vo EXAMPLE Figure 4.14 The noninverting op-amp configuration. RF RI Once again, we employ the ideal op-amp model conditions, that is, v- = v+ and i- = i+ . Using the fact that i- = 0 and v- = vin , the KCL equation at the negative terminal of the op-amp is vin vo - vin = RI RF or vin a vo 1 1 + b = RI RF RF Thus, vo RF = 1 + vin RI Note the similarity of this case to the inverting op-amp configuration in the previous example. We find that the gain in this configuration is also controlled by a simple resistor ratio but is not inverted; that is, the gain ratio is positive. SOLUTION 4.3 irwin04-156-188hr.qxd 166 9-07-2010 CHAPTER 4 14:17 Page 166 O P E R AT I O N A L A M P L I F I E R S The remaining examples, though slightly more complicated, are analyzed in exactly the same manner as those outlined above. EXAMPLE Gain error in an amplifier is defined as 4.4 GE = c actual gain - ideal gain d * 100% ideal gain We wish to show that for a standard noninverting configuration with finite gain A o , the gain error is GE = -100% 1 + Ao where = R1兾(R1 + R2). SOLUTION The standard noninverting configuration and its equivalent circuit are shown in Fig. 4.15a and b, respectively. The circuit equations for the network in Fig. 4.15b are vS = vin + v1 , vin = vo Ao R1 v = vo R1 + R2 o and v1 = The expression that relates the input and output is vS = vo c 1 + Ao 1 + d = vo c d Ao Ao and thus the actual gain is vo Ao = vS 1 + Ao Recall that the ideal gain for this circuit is (R1 + R2)兾R1 = 1兾. Therefore, the gain error is Ao 1 § 100% GE = £ 1 + A o 1兾 which, when simplified, yields GE = -100% 1 + Ao vS Figure 4.15 vo + Circuits used in Example 4.4. vS ± – R2 vo vin Aovin - v1 R2 R1 R1 (a) (b) ± – irwin04-156-188hr.qxd 9-07-2010 14:17 Page 167 SECTION 4.3 F U N D A M E N TA L O P - A M P C I R C U I T S Consider the op-amp circuit shown in Fig. 4.16. Let us determine an expression for the output voltage. R2 R1 v– + + v1 R3 v2 i– v+ i+ – ± R4 EXAMPLE 4.5 Figure 4.16 Differential amplifier operational amplifier circuit. + vo - - - 167 The node equation at the inverting terminal is SOLUTION v1 - vvo - v+ = iR1 R2 At the noninverting terminal KCL yields v2 - v+ v+ = + i+ R3 R4 However, i+ = i- = 0 and v+ = v- . Substituting these values into the two preceding equations yields v1 - vvo - v+ = 0 R1 R2 and v2 - vv = R3 R4 Solving these two equations for vo results in the expression vo = R2 R1 R4 R2 a1 + b v v R1 R2 R3 + R4 2 R1 1 Note that if R4 = R2 and R3 = R1 , the expression reduces to vo = R2 Av - v1 B R1 2 Therefore, this op-amp can be employed to subtract two input voltages. The circuit shown in Fig. 4.17a is a precision differential voltage-gain device. It is used to provide a single-ended input for an analog-to-digital converter. We wish to derive an expression for the output of the circuit in terms of the two inputs. To accomplish this, we draw the equivalent circuit shown in Fig. 4.17b. Recall that the voltage across the input terminals of the op-amp is approximately zero and the currents into the op-amp input terminals are approximately zero. Note that we can write node equations for node voltages v1 and v2 in terms of vo and va . Since we are interested in an expression for EXAMPLE 4.6 SOLUTION irwin04-156-188hr.qxd 168 9-07-2010 CHAPTER 4 14:17 Page 168 O P E R AT I O N A L A M P L I F I E R S vo in terms of the voltages v1 and v2 , we simply eliminate the va terms from the two node equations. The node equations are v1 - vo v1 - va v1 - v2 + + = 0 R2 R1 RG v2 - va v2 - v1 v2 + + = 0 R1 RG R2 Combining the two equations to eliminate va , and then writing vo in terms of v1 and v2 , yields vo = Av1 - v2 B a 1 + v1 Figure 4.17 Instrumentation amplifier circuit. ± – R2 2R2 + b R1 RG vo vo R2 R2 v1 v1 i1=0 R1 v2 ± – R1 va va RG RG R1 R1 v2 v2 i2=0 R2 R2 (a) (b) Learning Assessments ANSWER: Io = 8.4 mA. E4.1 Find Io in the network in Fig. E4.1. ± – Vo 12 k⍀ 12 V ± – 10 k⍀ 2 k⍀ Io Figure E4.1 irwin04-156-188hr.qxd 9-07-2010 14:17 Page 169 SECTION 4.3 F U N D A M E N TA L O P - A M P C I R C U I T S E4.2 Determine the gain of the op-amp circuit in Fig. E4.2. ANSWER: ± – 169 Vo R2 = 1 + . VS R1 + R2 VS ± – Vo R1 - Figure E4.2 E4.3 Determine both the gain and the output voltage of the op-amp configuration shown in Fig. E4.3. ± – 1 mV ± – ANSWER: Vo = 0.101 V; gain = 101. + Vo 100 k⍀ 1 k⍀ Figure E4.3 E4.4 Find I1, I2, I3, and I4 in Fig. E4.4. I1 5 k⍀ 1 mA I3 ± – Vo I2 5 k⍀ ANSWER: I1 = 0, I2 = 1.25 mA, I3 = -0.5 mA, I4 = 0.75 mA. I4 10 k⍀ 10 k⍀ Figure E4.4 E4.5 Find V0 in terms of V1 and V2 in Fig. E4.5. If V1 = V2 = 4 V, find V0. If the op-amp power supplies are 15 V and V2 = 2 V, what is the allowable range of V1? V1 5 k⍀ ANSWER: V0 = -2 V1 + 3.5 V2, 6 V, -4 V V1 11 V. 10 k⍀ 4 k⍀ V2 Vo – ± 4 k⍀ 10 k⍀ 7 k⍀ Figure E4.5 (continues on the next page) irwin04-156-188hr.qxd 170 9-07-2010 CHAPTER 4 14:17 Page 170 O P E R AT I O N A L A M P L I F I E R S E4.6 Find V0 and V3 in Fig. E4.6. ANSWER: V0 = -9 V, V3 = -4.8 V. 3 k⍀ 2 k⍀ V3 5 k⍀ 5V ± – + – Vo – ± 2.5 k⍀ 10 k⍀ 10 k⍀ 10 k⍀ 5 k⍀ Figure E4.6 E4.7 Find V0 in Fig. E4.7. ANSWER: V1 Vo ± – R3 R2 R1 ⎡⎛ R R ⎞ V0 = ⎢⎜ 3 + 3 + 1⎟ R R ⎠ 4 ⎣⎝ 2 ⎛ R2 ⎞ R3 ⎤ ⎜⎝1 + R ⎟⎠ − R ⎥ V1 1 2⎦ R4 Figure E4.7 EXAMPLE The two op-amp circuits shown in Fig. 4.18 produce an output given by the equation 4.7 where Vo = 8V1 - 4V2 1 V V1 2 V and 2 V V2 3 V We wish to determine (a) the range of Vo and (b) if both of the circuits will produce the full range of Vo given that the dc supplies are ;10 V. SOLUTION a. Given that Vo = 8 V1 - 4 V2 and the range for both V1 and V2 as 1 V V1 2 V and 2 V V2 3 V, we find that Vo max = 8(2) - 4(2) = 8 V and Vo min = 8(1) - 4(3) = -4 V and thus the range of Vo is -4 V to +8 V. b. Consider first the network in Fig. 4.18a. The signal at Vx , which can be derived using the network in Example 4.5, is given by the equation Vx = 2 V1 - V2 . Vx is a maximum when V1 = 2 V and V2 = 2 V, that is, Vx max = 2(2) - 2 = 2 V. The minimum value for Vx occurs when V1 = 1 V and V2 = 3 V, that is, Vx min = 2(1) - 3 = -1 V. Since both the max and min values are within the supply range of ; 10 V, the first op-amp in irwin04-156-188hr.qxd 9-07-2010 14:17 Page 171 SECTION 4.3 F U N D A M E N TA L O P - A M P C I R C U I T S Fig. 4.18a will not saturate. The output of the second op-amp in this circuit is given by the expression Vo = 4Vx . Therefore, the range of Vo is -4 V Vo 8 V. Since this range is also within the power supply voltages, the second op-amp will not saturate, and this circuit will produce the full range of Vo . Next, consider the network in Fig. 4.18b. The signal Vy = -8V1 and so the range of Vy is -16 V Vy -8 V and the range of Vy is outside the power supply limits. This circuit will saturate and fail to produce the full range of Vo . V1 Figure 4.18 Vx ± – ± – 10 k⍀ Circuits used in Example 4.7. Vo 30 k⍀ 10 k⍀ 10 k⍀ V2 (a) 80 k⍀ V1 10 k⍀ – ± V2 ± – Vy Vz 10 k⍀ 10 k⍀ 10 k⍀ – ± Vo 30 k⍀ 10 k⍀ (b) If you review the op-amp circuits presented in this chapter to this point, you will note one common characteristic of all circuits. The output is connected to the inverting input of the op-amp through a resistive network. This connection where a portion of the output voltage is fed back to the inverting input is referred to as negative feedback. Recall from the model of an ideal op-amp that the output voltage is proportional to the voltage difference between the input terminals. Feeding back the output voltage to the negative input terminal maintains this voltage difference near zero to allow linear operation of the op-amp. As a result, negative feedback is necessary for the proper operation of nearly all op-amp circuits. Our analysis of op-amp circuits is based on the assumption that the voltage difference at the input terminals is zero. Almost all op-amp circuits utilize negative feedback. However, positive feedback is utilized in oscillator circuits, the Schmitt trigger, and the comparator, which will be discussed in the following section. Let’s now consider the circuit in Fig. 4.19. This circuit is very similar to the circuit of Fig. 4.13a. However, there is one very important difference. In Fig. 4.19, resistor R2 is connected to the positive input terminal of the op-amp instead of the negative 171 irwin04-156-188hr.qxd 172 9-07-2010 CHAPTER 4 14:17 Page 172 O P E R AT I O N A L A M P L I F I E R S Figure 4.19 R2 Op-amp circuit with positive feedback. Vs R1 VCC Vo + VEE input. Connecting the output terminal to the positive input terminal results in positive feedback. As a result of the positive feedback, the output value of this op-amp circuit has two possible values VCC or VEE. Analysis of this circuit using the ideal op-amp model presented in this chapter does not predict this result. It is important to remember that the ideal op-amp model may only be utilized when negative feedback is present in the op-amp circuit. 4.4 Comparators A comparator, a variant of the op-amp, is designed to compare the noninverting and inverting input voltages. As shown in Fig. 4.20, when the noninverting input voltage is greater, the output goes as high as possible, at or near Vcc. On the other hand, if the inverting input voltage is greater, the output goes as low as possible, at or near VEE. Of course, an ideal op-amp can do the same thing, that is, swing the output voltage as far as possible. However, op-amps are not designed to operate with the outputs saturated, whereas comparators are. As a result, comparators are faster and less expensive than op-amps. We will present two very different quad comparators in this text, National Semiconductor’s LM339 and Maxim’s MAX917. Note that the LM339 requires a resistor, called a pull-up resistor, connected between the output pin and Vcc. The salient features of these products are listed in Table 4.4. From Table 4.4, it is easy to surmise that the LM339 is a general-purpose comparator, whereas the MAX917 is intended for low-power applications such as hand-held products. (a) An ideal comparator and (b) its transfer curve. VCC vin+ - V+ ± – V– ± – ± – + VEE Vo - Output voltage Figure 4.20 VCC 0 VEE –1.5 –1 –0.5 0 0.5 1 1.5 Input voltage (V+-V–) (a) (b) A common comparator application is the zero-crossing detector, shown in Fig. 4.21a using a LM339 with ;5 V supplies. As seen in Fig. 4.21b, when VS is positive, Vo should be near +5 V and when VS is negative, Vo should be near -5 V. The output changes value on every zero crossing! TABLE 4.4 PRODUCT LM339 MAX919 A listing of some of the features of the LM339 and MAX917 comparators MIN. SUPPLY MAX. SUPPLY SUPPLY CURRENT MAX. OUTPUT CURRENT TYPICAL Rpull - up 2V 2.8 V 36 V 5.5 V 3 mA 0.8 A 50 mA 8 mA 3 k NA irwin04-156-188hr.qxd 9-07-2010 14:17 Page 173 SECTION 4.5 Input Output 6 3 k⍀ 4 –5 V VS ± – I/O voltages (V) +5 V ± – A P P L I C AT I O N E X A M P L E S + Vo - 2 173 Figure 4.21 (a) A zero-crossing detector and (b) the corresponding input/ output waveforms. 0 –2 –4 –6 Time (a) (b) At this point, we have a new element, the op-amp, which we can effectively employ in both applications and circuit design. This device is an extremely useful element that vastly expands our capability in these areas. Because of its ubiquitous nature, the addition of the op-amp to our repertoire of circuit elements permits us to deal with a wide spectrum of practical circuits. Thus, we will employ it here, and also use it throughout this text. In a light meter, a sensor produces a current proportional to the intensity of the incident radiation. We wish to obtain a voltage proportional to the light’s intensity using the circuit in Fig. 4.22. Thus, we select a value of R that will produce an output voltage of 1 V for each 10 A of sensor current. Assume that the sensor has zero resistance. 4.5 Application Examples APPLICATION EXAMPLE 4.8 • Figure 4.22 R Light intensity to voltage converter. – ± I + Incident light Light sensor Vo - SOLUTION Applying KCL at the op-amp input, Since Vo兾I is 105, I = Vo兾R R = 100 k The circuit in Fig. 4.23 is an electronic ammeter. It operates as follows: the unknown current, I, through RI produces a voltage, VI . VI is amplified by the op-amp to produce a voltage, Vo , which is proportional to I. The output voltage is measured with a simple voltmeter. We want to find the value of R2 such that 10 V appears at Vo for each milliamp of unknown current. APPLICATION EXAMPLE 4.9 • irwin04-156-188hr.qxd 174 9-07-2010 CHAPTER 4 14:17 Page 174 O P E R AT I O N A L A M P L I F I E R S SOLUTION Since the current into the op-amp + terminal is zero, the relationship between VI and I is VI = IRI The relationship between the input and output voltages is Vo = VI a 1 + R2 b R1 or, solving the equation for Vo兾I, we obtain Vo R2 b = RI a 1 + I R1 Using the required ratio Vo兾I of 104 and resistor values from Fig. 4.23, we can find that R2 = 9 k Figure 4.23 ± – I Electronic ammeter. Unknown current Voltmeter + VI RI=1 k⍀ Vo - R2 SOLUTION - + R1=1 k⍀ APPLICATION EXAMPLE 4.10 + - Let us return to the dc motor control example in Chapter 3 (Example 3.22). We want to define the form of the power amplifier that reads the speed control signal, Vspeed, and outputs the dc motor voltage with sufficient current to drive the motor as shown in Fig. 4.24. Let us make our selection under the condition that the total power dissipation in the amplifier should not exceed 100 mW. From Table 4.1 we find that the only op-amp with sufficient output voltage—that is, a maximum output voltage of (4)(5) = 20 V—for this application is the PA03 from APEX. Since the required gain is +4, we can employ the standard noninverting amplifier configuration shown in Fig. 4.25. If the PA03 is assumed to be ideal, then VM = Vspeed c 1 + RB d = 4Vspeed RA There are, of course, an infinite number of solutions that will satisfy this equation. Figure 4.24 The dc motor example from Chapter 3. 5V Rpot R1 ␣=1 R2 ␣=0 + Vspeed – Power amp VM/Vspeed=4 + VM – dc motor irwin04-156-188hr.qxd 9-07-2010 14:17 Page 175 SECTION 4.5 A P P L I C AT I O N E X A M P L E S ± – Figure 4.25 R1 5V The power amplifier configuration using the PA03 op-amp. + + VM RB R2 Vspeed – RA – 175 In order to select reasonable values, we should consider the possibility of high currents in RA and RB when VM is at its peak value of 20 V. Assuming that Rin for the PA03 is much greater than RA , the currents in RB and RA essentially determine the total power dissipated. The total power dissipated in RA and RB is Ptotal = V2M 202 400 = RA + RB RA + RB RA + RB Since the total power should not exceed 100 mW, we can use 1兾4 W resistors—an inexpensive industry standard—with room to spare. With this power specification, we find that RA + RB = V2M 400 = = 4000 Ptotal 0.1 Also, since 1 + RB = 4 RA then RB = 3 RA . Combining this result with the power specification yields RA = 1 k and RB = 3 k. Both are standard 5% tolerance values. An instrumentation amplifier of the form shown in Fig. 4.26 has been suggested. This amplifier should have high-input resistance, achieve a voltage gain Vo兾(V1 - V2) of 10, employ the MAX4240 op-amp listed in Table 4.1, and operate from two 1.5 V AA cell batteries in series. Let us analyze this circuit, select the resistor values, and explore the validity of this configuration. Differential amplifier Figure 4.26 3V + V1 – 3V RA RB=RA A 3V Vx ± – ± – R 1.5 V B 1.5 V + V2 – R2 – ± RA Vy 3V + Vo – R1 RB=RA APPLICATION EXAMPLE 4.11 An instrumentation amplifier using the MAX4240 op-amp. • irwin04-156-188hr.qxd 176 9-07-2010 CHAPTER 4 14:17 Page 176 O P E R AT I O N A L A M P L I F I E R S SOLUTION As indicated, the op-amp on the right side of the circuit is connected in the traditional differential amplifier configuration. Example 4.5 indicates that the voltage gain for this portion of the network is Vo = (Vx - Vy) c RB d RA And if RA = RB , the equation reduces to Vo = Vx - Vy If we can find a relationship between V1 , V2 , and Vx and Vy , then an expression for the overall voltage can be written. Applying KCL at node A yields V1 - V2 Vx - V1 = R R1 or Vx = V1 c 1 + R1 R1 d - V2 c d R R In a similar manner, at node B we obtain V2 - Vy V1 - V2 = R R2 or Vy = -V1 c R2 R2 d + V2 c 1 + d R R By combining these equations, the output voltage can be expressed as Vo = Vx - Vy = V1 c 1 + R1 R2 R1 R2 d - V2 c d + V1 c d - V2 c 1 + d R R R R If the resistors are selected such that R1 = R2 , then the voltage gain is Vo 2R1 = 1 + V1 - V2 R For a gain of +10, we set R1 = 4.5 R. To maintain low power, we will use fairly large values for these resistors. We somewhat arbitrarily choose R = 100 k and R1 = R2 = 450 k. We can use 100 k resistors in the differential amplifier stage as well. Note that the voltage gain of the instrumentation amplifier is essentially the same as that of a generic differential amplifier. So why add the cost of two more op-amps? In this configuration the inputs V1 and V2 are directly connected to op-amp input terminals; therefore, the input resistance of the intrumentation amplifier is extremely large. From Table 4.1 we see that Rin for the MAX4240 is 45 M. This is not the case in the traditional differential amplifier where the external resistor can significantly decrease the input resistance. 4.6 Design Examples DESIGN EXAMPLE 4.12 We are asked to construct an amplifier that will reduce a very large input voltage (i.e., Vin ranges between ;680 V) to a small output voltage in the range < 5 V. Using only two resistors, we wish to design the best possible amplifier. irwin04-156-188hr.qxd 9-07-2010 14:17 Page 177 SECTION 4.6 DESIGN EXAMPLES Since we must reduce +680 V to -5 V, the use of an inverting amplifier seems to be appropriate. The input/output relationship for the circuit shown in Fig. 4.27 is 177 SOLUTION Vo R2 = Vin R1 Since the circuit must reduce the voltage, R1 must be much larger than R2 . By trial and error, one excellent choice for the resistor pair, selected from the standard Table 2.1, is R1 = 27 k and R2 = 200 . For Vin = 680 V, the resulting output voltage is 5.037 V, resulting in a percent error of only 0.74%. Figure 4.27 R2 I R1 A standard inverting amplifier stage. – ± + Vin ± – Vo - There is a requirement to design a noninverting op-amp configuration with two resistors under the following conditions: the gain must be +10, the input range is ;2 V, and the total power consumed by the resistors must be less than 100 mW. DESIGN EXAMPLE 4.13 SOLUTION For the standard noninverting configuration in Fig. 4.28a, the gain is Vo R2 = 1 + Vin R1 For a gain of 10, we find R2兾R1 = 9. If R1 = 3 k and R2 = 27 k, then the gain requirement is met exactly. Obviously, a number of other choices can be made, from the standard Table 2.1, with a 3兾27 ratio. The power limitation can be formalized by referring to Fig. 4.28b where the maximum input voltage (2 V) is applied. The total power dissipated by the resistors is PR = (20 - 2)2 22 4 324 + = + 6 0.1 R1 R2 R1 9R1 The minimum value for R1 is 400 . ± – Vin ± – Vo R2 R1 (a) ± – + - 2V ± – Figure 4.28 + Vo=20 V 2V R2 R1 (b) - The noninverting op-amp configuration employed in Example 4.13. • irwin04-156-188hr.qxd 178 9-07-2010 CHAPTER 4 14:17 Page 178 O P E R AT I O N A L A M P L I F I E R S DESIGN EXAMPLE 4.14 We wish to design a weighted-summer circuit that will produce the output Vo = -0.9V1 - 0.1V2 The design specifications call for use of one op-amp and no more than three resistors. Furthermore, we wish to minimize power while using resistors no larger than 10 k. SOLUTION A standard weighted-summer configuration is shown in Fig. 4.29. Our problem is reduced to finding values for the three resistors in the network. Using KCL, we can write I1 + I2 = - Vo R where I1 = V1 R1 and I2 = V2 R2 Combining these relationships yields Vo = - c R R d V1 - c d V2 R1 R2 Therefore, we require R = 0.9 and R1 R = 0.1 R2 From these requirements, we see that the largest resistor is R2 and that R is the smallest. Also, note that the R兾R1 ratio can be expressed as 27兾30. Finally, to minimize power, we should use the largest possible resistor values. Based on this information, the best resistor values are R = 270 , R1 = 300 , and R2 = 2.7 k, which yield the desired performance exactly. R1 I1 Figure 4.29 A standard weightedsummer configuration. I2 V1 ± – V2 ± – R2 R – ± + Vo - DESIGN EXAMPLE 4.15 SOLUTION In Example 2.36, a 250- resistor was used to convert a current in the 4- to 20-mA range to a voltage such that a 20-mA input produced a 5-V output. In this case, the minimum current (4 mA) produces a resistor voltage of 1 V. Unfortunately, many control systems operate on a 0- to 5-V range rather than a 1- to 5-V range. Let us design a new converter that will output 0 V at 4 mA and 5 V at 20 mA. The simple resistor circuit we designed in Example 2.36 is a good start. However, the voltage span is only 4 V rather than the required 5 V, and the minimum value is not zero. These irwin04-156-188hr.qxd 9-07-2010 14:17 Page 179 SUMMARY 179 facts imply that a new resistor value is needed and the output voltage should be shifted down so that the minimum is zero. We begin by computing the necessary resistor value: R = Vmax - Vmin 5 - 0 = = 312.5 Imax - Imin 0.02 - 0.004 The resistor voltage will now range from (0.004)(312.5) to (0.02)(312.5), or 1.25 to 6.25 V. We must now design a circuit that shifts these voltage levels so that the range is 0 to 5 V. One possible option for the level shifter circuit is the differential amplifier shown in Fig. 4.30. Recall that the output voltage of this device is Vo = (VI - Vshift) R2 R1 Since we have already chosen R for a voltage span of 5 V, the gain of the amplifier should be 1 (i.e., R1 = R2). Clearly, the value of the required shift voltage is 1.25 V. However, we can verify this value by inserting the minimum values into this last equation 0 = [(312.5)(0.004) - Vshift] R2 R1 and find Vshift = (312.5)(0.004) = 1.25 V There is one caveat to this design. We don’t want the converter resistor, R, to affect the differential amplifier, or vice versa. This means that the vast majority of the 4–20 mA current should flow entirely through R and not through the differential amplifier resistors. If we choose R1 and R2 W R, this requirement will be met. Therefore, we might select R1 = R2 = 100 k so that their resistance values are more than 300 times that of R. Figure 4.30 Differential amplifier with shifter A 4–20 mA to 0–5 V converter circuit. R1 ± – + R VI 4–20 mA - • R2 R1 + R2 Vo ± Vshift – - SUMMARY ■ Op-amps are characterized by High-input resistance Low-output resistance Very high gain ■ The ideal op-amp is modeled using i+ = i- = 0 v+ = v- ■ Op-amp problems are typically analyzed by writing node equations at the op-amp input terminals ■ The output of a comparator is dependent on the difference in voltage at the input terminals irwin04-156-188hr.qxd 180 9-07-2010 CHAPTER 4 14:17 Page 180 O P E R AT I O N A L A M P L I F I E R S PROBLEMS • 4.1 An amplifier has a gain of 15 and the input waveform shown in Fig. P4.1. Draw the output waveform. vo(V) 3 vin(mV) 2 150 1 100 0 50 25 50 75 100 125 t (ms) –1 0 0.5 1 1.5 2.0 t (s) –2 –50 –3 –100 Figure P4.3 –150 4.4 For an ideal op-amp, the voltage gain and input resistance Figure P4.1 are infinite while the output resistance is zero. What are the consequences for 4.2 An amplifier has a gain of -5 and the output waveform shown in Fig. P4.2. Sketch the input waveform. (a) the op-amp’s input voltage? (b) the op-amp’s input currents? (c) the op-amp’s output current? vo(V) 4.5 Revisit your answers in Problem 4.4 under the following 12 nonideal scenarios. 10 (a) Rin = q, Rout = 0, A o Z q. 6 (b) Rin = q, Rout 7 0, A o = q. 5 (c) Rin Z q, Rout = 0, A o = q. 4.6 Revisit the exact analysis of the inverting configuration in 4 Section 4.3. 2 0 (a) Find an expression for the gain if Rin = q, Rout = 0, 1 2 3 4 5 6 7 8 9 t (ms) A o Z q. (b) Plot the ratio of the gain in (a) to the ideal gain versus –2 A o for 1 A o 1000for an ideal gain of -10. (c) From your plot, does the actual gain approach the ideal value as A o increases or decreases? (d) From your plot, what is the minimum value of A o if the actual gain is within 5% of the ideal case? –4 –6 –8 –10 4.7 Revisit the exact analysis of the inverting amplifier in –12 Figure P4.2 4.3 An op-amp based amplifier has supply voltages of ;5 V and a gain of 20. (a) Sketch the input waveform from the output waveform in Fig. P4.3. (b) Double the amplitude of your results in (a) and sketch the new output waveform. Section 4.3. (a) Find an expression for the voltage gain if Rin Z q, Rout = 0, A o Z q. (b) For R2 = 27 k and R1 = 3 k, plot the ratio of the actual gain to the ideal gain for A o = 1000 and 1 k Rin 100 k. (c) From your plot, does the ratio approach unity as Rin increases or decreases? (d) From your plot in (b), what is the minimum value of Rin if the gain ratio is to be at least 0.98? irwin04-156-188hr.qxd 9-07-2010 14:17 Page 181 PROBLEMS 4.8 An op-amp based amplifier has ;18 V supplies and a gain 181 4.13 Assuming an ideal op-amp in Fig. P4.13, determine the output voltage Vo. of -80. Over what input range is the amplifier linear? 4 k⍀ 4.9 Assuming an ideal op-amp, determine the voltage gain of the 1 k⍀ circuit in Fig. P4.9. – ± 20 k⍀ 6V 1 k⍀ – ± ± – + ± – 2V Vo 2 k⍀ - + Figure P4.13 vo ± v – 1 4.14 Determine the gain of the amplifier in Fig. P4.14. What - is the value of Io ? Figure P4.9 Io ± – Vin 4.10 Assuming an ideal op-amp, determine the voltage gain of Vo the circuit in Fig. P4.10. R2 + - 4 k⍀ + 2 k⍀ v1 ± – R2=20 kΩ R1=3.3 kΩ Vin=2 V R1 vo 200 k⍀ - Figure P4.14 4.15 For the amplifier in Fig. P4.15, find the gain and Io . Figure P4.10 value of RX that will produce a voltage gain of 26. 4 k⍀ v1 Vo R2 ± – + Io ± – 4.11 Assuming an ideal op-amp in Fig. P4.11, determine the R2=20 kΩ R1=3.3 kΩ VS=2 V + vo RX - - R1 Figure P4.11 4.12 Assuming an ideal op-amp, find the voltage gain of the network in Fig. P4.12. VS Figure P4.15 4.16 Using the ideal op-amp assumptions, determine the 18 k⍀ values of Vo and I1 in Fig. P4.16. 9 k⍀ 100 ⍀ + 150 ⍀ 11 V – ± ± – + v1 vo - - Figure P4.12 I1 10 k⍀ 1 k⍀ Figure P4.16 Vo irwin04-156-188hr.qxd 182 9-07-2010 CHAPTER 4 14:17 Page 182 O P E R AT I O N A L A M P L I F I E R S 4.17 Using the ideal op-amp assumptions, determine I1 , I2 , and I3 in Fig. P4.17. VS ± – Vo I3 R2 I1 R1 I2 1 mA R2 R1 – ± RL=10 kΩ R1=100 kΩ RL Vo Figure P4.20 Figure P4.17 4.18 In a useful application, the amplifier drives a load. The 4.21 For the circuit in Fig. P4.21, circuit in Fig. P4.18 models this scenario. (a) find Vo in terms of V1 and V2 . (a) Sketch the gain Vo兾VS for 10 RL q. (b) If V1 = 2 V and V2 = 6 V, find Vo . (b) Sketch Io for 10 RL q if VS = 0.1 V. (c) If the op-amp supplies are ;12 V, and V1 = 4 V, (c) Repeat (b) if VS = 1.0 V. (d) What is the minimum value of RL if ∑Io∑ must be less than 100 mA for ∑VS∑ 6 0.5 V? (e) What is the current IS if RL is 100 ? Repeat for RL = 10 k. IS ± – VS Io Vo what is the allowable range of V2 ? V1 ± – V2 Vo 2 k⍀ 2 k⍀ 1 k⍀ R2=27 kΩ R1=3 kΩ R2 Figure P4.21 RL R1 4.22 Find Vo in the circuit in Fig. P4.22, assuming that the op-amp is ideal. Figure P4.18 4.19 The op-amp in the amplifier in Fig. P4.19 operates with 1 k⍀ ; 15 V supplies and can output no more than 200 mA. What is the maximum gain allowable for the amplifier if the maximum value of VS is 1 V? VS ± – R2 Vo RL ± – R1+R2=10 kΩ Figure P4.19 4.23 The network in Fig. P4.23 is a current-to-voltage converter or transconductance amplifier. Find vo兾iS for this network. 1⍀ VS is 2 V and the op-amp can deliver no more than 100 mA. allowable value of R2 ? + iS (b) Repeat for ; 3 V supplies. + vo - (c) Discuss the impact of the supplies on the maximum allowable gain. Vo Figure P4.22 4.20 For the amplifier in Fig. P4.20, the maximum value of (a) If ;10 V supplies are used, what is the maximum + 1V - 50 ⍀ R1 5 k⍀ + Figure P4.23 irwin04-156-188hr.qxd 9-07-2010 14:17 Page 183 PROBLEMS 4.24 Calculate the transfer function io兾v1 for the network R3 V1 shown in Fig. P4.24. v1 ± – 183 Vo io ± – R1 R4 R2 RF RI V2 Figure P4.28 Figure P4.24 4.25 Determine the relationship between v1 and io in the 4.29 Find Vo in the network in Fig. P4.29. circuit shown in Fig. P4.25. 4⍀ RF RI v1 R1 1⍀ + io 5V ± – 4V + + ± – Vo - Figure P4.25 Figure P4.29 4.26 Find Vo in the network in Fig. P4.26 and explain what effect R1 has on the output. 4.30 Find the voltage gain of the op-amp circuit shown in Fig. P4.30. 10 ⍀ 20 k⍀ 2⍀ + 2V ± – 10 ⍀ + Vo R1 V1 ± – 80 k⍀ ± – + 24 k⍀ Vo 1 k⍀ - - Figure P4.26 Figure P4.30 4.27 Determine the expression for vo in the network in Fig. P4.27. 4.31 Determine the relationship between vo and vin in the R2 vA circuit in Fig. P4.31. R1 + vB vo vin R1 Figure P4.27 R2 RI 4.28 Show that the output of the circuit in Fig. P4.28 is Vo = c 1 + ± – R2 R2 dV V R1 1 R1 2 Figure P4.31 RF vo irwin04-156-188hr.qxd 184 9-07-2010 CHAPTER 4 14:17 Page 184 O P E R AT I O N A L A M P L I F I E R S 4.32 In the network in Fig. P4.32, derive the expression for vo in terms of the inputs v1 and v2 . the inverting-summer circuit shown in Fig. P4.36. + vo R2 RI RF R1 v1 RF R1 v1 v2 4.36 Determine the expression for the output voltage, vo , of v2 R2 v3 R3 + vo Figure P4.36 Figure P4.32 4.37 Determine the output voltage, vo , of the noninverting averaging circuit shown in Fig. P4.37. 4.33 For the circuit in Fig. P4.33, find the value of R1 that produces a voltage gain of 10. V1 ± – ± – + 18 k⍀ Vo R1 R1 v1 v2 R2 v3 R3 ± – RF RI - vo Figure P4.37 Figure P4.33 4.38 Find the input/output relationship for the current amplifier shown in Fig. P4.38. 4.34 Find Vo in the circuit in Fig. P4.34. RF 40 k⍀ iin + 5 k⍀ + ± – 5V ± – Vo 20 k⍀ RI - Figure P4.38 Figure P4.34 4.39 Find Vo in the circuit in Fig. P4.39. 80 k⍀ 4.35 Find Vo in the circuit in Fig. P4.35. 10 k⍀ 40 k⍀ 100 k⍀ ± – 10 k⍀ 20 k⍀ 9V + 30 k⍀ 10 V ± – Figure P4.35 - 12 V + 40 k⍀ ± – 20 k⍀ 6V + Vo io RL + 5 k⍀ 4V ± – ± – + + 5V + Vo - - Figure P4.39 20 k⍀ 40 k⍀ irwin04-156-188hr.qxd 9-07-2010 14:17 Page 185 PROBLEMS 4.42 Find vo in the circuit in Fig. P4.42. 4.40 Find vo in the circuit in Fig. P4.40. R2 v1 R2 R1 ± – R1 – ± v2 185 R3 R4 vo R1 v1 R2 – ± vo Figure P4.42 4.43 Find the output voltage, vo , in the circuit in Fig. P4.43. Figure P4.40 4.41 Find the expression for vo in the differential amplifier R4 v1 circuit shown in Fig. P4.41. + - RF R1 v1 – ± – ± R1 R2 R1 R2 R1 R1 v2 vo – ± R3 – ± v2 R3 R4 Figure P4.41 Figure P4.43 4.44 The electronic ammeter in Example 4.9 has been modified and is shown in Fig. P4.44. The selector switch allows the user to change the range of the meter. Using values for R1 and R2 from Example 4.9, find the values of RA and RB that will yield a 10-V output when the current being measured is 100 mA and 10 mA, respectively. ± – I Unknown current RA RB RC=1 k⍀ Voltmeter + Vo R2=9 k⍀ Selector switch Figure P4.44 R1=1 k⍀ + - irwin04-156-188hr.qxd 186 9-07-2010 CHAPTER 4 14:17 Page 186 O P E R AT I O N A L A M P L I F I E R S 4.45 Given a box of 10-k resistors and an op-amp, design a circuit that will have an output voltage of 4.51 Design an op-amp-based circuit to produce the function Vo = -2V1 - 4V2 4.46 Design an op-amp circuit that has a gain of -50 using resistors no smaller than 1 k. Vo = 5 V1 - 4 V2 4.52 Design an op-amp-based circuit to produce the function Vo = 5 V1 - 7 V2 4.47 Design a two-stage op-amp network that has a gain of -50,000 while drawing no current into its input terminal. Use no resistors smaller than 1 k. 4.53 Show that the circuit in Fig. P4.53 can produce the output Vo = K1 V1 - K2 V2 4.48 Design an op-amp circuit that has the following input/output relationship: only for 0 K1 K2 + 1. Vo = -5 V1 + 0.5 V2 V1 R1 4.49 A voltage waveform with a maximum value of Vo R2 200 mV must be amplified to a maximum of 10 V and inverted. However, the circuit that produces the waveform can provide no more than 100 A. Design the required amplifier. R4 R3 4.50 An amplifier with a gain of ; 1% is needed. Using resistor values from Table 2.1, design the amplifier. Use as few resistors as possible. ± – V2 Figure P4.53 irwin04-156-188hr.qxd 9-07-2010 14:17 Page 187 TYPICAL PROBLEMS FOUND ON THE FE EXAM 187 TYPICAL PROBLEMS FOUND ON THE FE EXAM • 4PFE-1 Given the summing amplifier shown in Fig. 4PFE-1, select the values of R2 that will produce an output voltage of –3 V. a. 4.42 k c. 3.6 k b. 6.33 k d. 5.14 k R2 4 k⍀ – ± 4V ± – + 12 k⍀ – ± Vo 2V - Figure 4PFE-1 4PFE-2 Determine the output voltage Vo of the summing op-amp circuit shown in Fig. 4PFE-2. a. 6 V c. 9 V b. 18 V d. 10 V 18 k⍀ 36 k⍀ 6 k⍀ 6 k⍀ – ± 2V ± – – ± 12 k⍀ 12 k⍀ – ± 1V ± – Vo 3V - Figure 4PFE-2 4PFE-3 What is the output voltage Vo in Fig. 4PFE-3? a. 5 V c. 4 V b. 6 V d. 7 V 2⍀ + 3⍀ 6V ± – 4⍀ 2V ± – Figure 4PFE-3 + + Vo - irwin04-156-188hr.qxd 188 9-07-2010 CHAPTER 4 14:17 Page 188 O P E R AT I O N A L A M P L I F I E R S 4PFE-4 What value of Rf in the op-amp circuit of Fig. 4PFE-4 4PFE-5 What is the voltage Vo in the circuit in Fig. 4PFE-5? is required to produce a voltage gain of 50? a. 3 V a. 135 k b. 6 V b. 210 k c. 8 V c. 180 k d. 5 V d. 245 k 6 k⍀ Vin + - + - + Vo 5V ± – - 8 k⍀ Rf R1 = 5 k⍀ + 2 k⍀ 1 k⍀ Vo 2 k⍀ - Figure 4PFE-4 Figure 4PFE-5 irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 189 C H A PT E R ADDITIONAL ANALYSIS TECHNIQUES 5 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Understand the concepts of linearity and equivalence ■ Know how to analyze electric circuits using the principle of superposition ■ Be able to calculate a Thévenin equivalent circuit for a linear circuit ■ Be able to calculate a Norton equivalent circuit for a linear circuit ■ Understand when and how to use a source transformation ■ Be able to use the maximum power transfer theorem Courtesy egdigital/iStockphoto M Monitoring Devices Devices that monitor traffic flow on inter- Measuring and regulating current flow in a circuit is as state roadways help reduce congestion near cities. For improved important as ensuring traffic flows freely. In this chapter, we traffic control, monitoring devices flash your speed as you use the superposition concept to calculate currents flowing in approach, warning you when you are over the limit and persuad- an electric circuit due to multiple sources. We form a simple ing you to slow down. Such devices are useful in construction equivalent circuit for parts of circuits that remain fixed, zones where reduced speeds are required to protect workers. enabling a focus only on components to be changed for differ- Electric circuits have similar tools: current-monitoring ent operating conditions. Once you’ve mastered these con- devices that detect excessive currents and disconnect compo- cepts you’ll be better prepared to design circuits that keep the nents that would be damaged if power limitations are exceed- current flowing under the legal limit. ed. These circuit breakers in homes or office buildings have cutoff features that can be reset to restore service. 189 irwin05-189-244hr2.qxd 190 22-07-2010 CHAPTER 5 5.1 Introduction 9:47 Page 190 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Before introducing additional analysis techniques, let us review some of the topics we have used either explicitly or implicitly in our analyses thus far. EQ U I VA L E N C E Table 5.1 is a short compendium of some of the equivalent circuits that have been employed in our analyses. This listing serves as a quick review as we begin to look at other techniques that can be used to find a specific voltage or current somewhere in a network and provide additional insight into the network’s operation. In addition to the forms listed in the table, it is important to note that a series connection of current sources or a parallel connection of voltage sources is forbidden unless the sources are pointing in the same direction and have exactly the same values. TABLE 5.1 Equivalent circuit forms R1 R1+R2 R2 R1 R1 R2 — R1+R2 R2 V1 ± – ± V -V 1 2 – V2 + I1 I2 I1-I2 + VS ± – R Vo=VS R Io=IS IS - L I N E A R I TY All the circuits we have analyzed thus far have been linear circuits, which are described by a set of linear algebraic equations. Most of the circuits we will analyze in the remainder of the book will also be linear circuits, and any deviation from this type of network will be specifically identified as such. Linearity requires both additivity and homogeneity (scaling). It can be shown that the circuits that we are examining satisfy this important property. The following example illustrates one way in which this property can be used. irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 191 SECTION 5.1 INTRODUCTION 191 EXAMPLE For the circuit shown in Fig. 5.1, we wish to determine the output voltage Vout˚. However, rather than approach the problem in a straightforward manner and calculate Io˚, then I1˚, then I2˚, and so on, we will use linearity and simply assume that the output voltage is Vout = 1 V. This assumption will yield a value for the source voltage. We will then use the actual value of the source voltage and linearity to compute the actual value of Vout˚. 5.1 SOLUTION If we assume that Vout = V2 = 1 V, then V2 = 0.5 mA 2k I2 = V1 can then be calculated as V1 = 4kI2 + V2 = 3V Hence, I1 = V1 = 1 mA 3k Now, applying KCL, Io = I1 + I2 = 1.5 mA Then Vo = 2kIo + V1 = 6V Therefore, the assumption that Vout = 1 V produced a source voltage of 6 V. However, since the actual source voltage is 12 V, the actual output voltage is 1 V(12兾6) = 2 V. Vo Io V1 I2 V2 2 k⍀ 12 V Figure 5.1 + 4 k⍀ ± – 3 k⍀ I1 2 k⍀ I2 Circuit used in Example 5.1. Vout - Learning Assessment E5.1 Use linearity and the assumption that Io = 1 mA to compute the correct current Io in the circuit in Fig. E5.1 if I = 6 mA. 4 k⍀ 2 k⍀ 8 k⍀ 6 k⍀ 3 k⍀ Io I Figure E5.1 E5.2 Find V0 in Fig. E5.2 using linearity and the assumption that Vo 1 V. 4 k⍀ 4 k⍀ 2 k⍀ + 12 k⍀ 20 V + – 8 k⍀ 6 k⍀ 4 k⍀ V0 – Figure E5.2 ANSWER: Io = 3 mA. 3 k⍀ ANSWER: Vo 5/3 V. irwin05-189-244hr2.qxd 192 22-07-2010 CHAPTER 5 9:47 Page 192 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.2 Superposition EXAMPLE To provide motivation for this subject, let us examine the simple circuit of Fig. 5.2a, in which two sources contribute to the current in the network. The actual values of the sources are left unspecified so that we can examine the concept of superposition. 5.2 SOLUTION The mesh equations for this network are 6ki1(t) - 3ki2(t) = v1(t) -3ki1(t) + 9ki2(t) = -v2(t) Solving these equations for i1(t) yields v1(t) i1(t) = 5k - v2(t) 15k In other words, the current i1(t) has a component due to v1(t) and a component due to v2(t). In view of the fact that i1(t) has two components, one due to each independent source, it would be interesting to examine what each source acting alone would contribute to i1(t). For v1(t) to act alone, v2(t) must be zero. As we pointed out in Chapter 2, v2(t)=0 means that the source v2(t) is replaced with a short circuit. Therefore, to determine the value of i1(t) due to v1(t) only, we employ the circuit in Fig. 5.2b and refer to this value of i1(t) as i1œ (t). i 1œ (t) = v1(t) v1(t) = (3k)(6k) 5k 3k + 3k + 6k Let us now determine the value of i1(t) due to v2(t) acting alone and refer to this value as i1fl(t). Using the network in Fig. 5.2c, i 2fl(t) = - v2(t) -2v2(t) = (3k)(3k) 15k 6k + 3k + 3k Then, using current division, we obtain Figure 5.2 i 1fl(t) = Circuits used to illustrate superposition. 3 k⍀ 6 k⍀ i'1(t) -2v2(t) 15k 3 k⍀ a -v2(t) 3k b = 3k + 3k 15k 6 k⍀ i"1(t) 3 k⍀ 6 k⍀ i"2(t) 3 k⍀ v1(t) ± – i1(t) i2(t) (a) ± v2(t) v1(t) ± – – 3 k⍀ 3 k⍀ (b) ± v2(t) – (c) Now, if we add the values of i1œ (t) and i1fl(t), we obtain the value computed directly; that is, i1(t) = i1œ (t) + i1fl(t) = v1(t) 5k - v2(t) 15k Note that we have superposed the value of i1œ (t) on i1fl(t), or vice versa, to determine the unknown current. irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 193 SECTION 5.2 SUPERPOSITION 193 What we have demonstrated in Example 5.2 is true in general for linear circuits and is a direct result of the property of linearity. The principle of superposition, which provides us with this ability to reduce a complicated problem to several easier problems—each containing only a single independent source—states that In any linear circuit containing multiple independent sources, the current or voltage at any point in the network may be calculated as the algebraic sum of the individual contributions of each source acting alone. When determining the contribution due to an independent source, any remaining voltage sources are made zero by replacing them with short circuits, and any remaining current sources are made zero by replacing them with open circuits. Although superposition can be used in linear networks containing dependent sources, it is not useful in this case since the dependent source is never made zero. As the previous example indicates, superposition provides some insight in determining the contribution of each source to the variable under investigation. We will now demonstrate superposition with two examples and then provide a problemsolving strategy for the use of this technique. For purposes of comparison, we will also solve the networks using both node and loop analyses. Furthermore, we will employ these same networks when demonstrating subsequent techniques, if applicable. EXAMPLE Let us use superposition to find Vo in the circuit in Fig. 5.3a. 2 k⍀ 1 k⍀ + 2 mA 2 k⍀ + 3V Vo 6 k⍀ 1 k⍀ Io 6 k⍀ 2 mA (a) (b) 2 k⍀ + + 3V 1 k⍀ V'o – – 2 k⍀ V"o 6 k⍀ 1 k⍀ 3V + I1 I2 + 6 k⍀ Vo 2 mA – (c) – Figure 5.3 Circuits used in Example 5.3. (d) The contribution of the 2-mA source to the output voltage is found from the network in Fig. 5.3b, using current division and 5.3 + Io = A2 * 10-3 B a 1k + 2k 2 b = mA 1k + 2k + 6k 3 Voœ = Io(6k) = 4 V The contribution of the 3-V source to the output voltage is found from the circuit in Fig. 5.3c. Using voltage division, Vofl = 3 a 6k b 1k + 2k + 6k = 2V Therefore, Vo = Voœ + Vofl = 6 V Although we used two separate circuits to solve the problem, both were very simple. SOLUTION irwin05-189-244hr2.qxd 194 22-07-2010 CHAPTER 5 9:47 Page 194 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S If we use nodal analysis and Fig. 5.3a to find Vo and recognize that the 3-V source and its connecting nodes form a supernode, Vo can be found from the node equation Vo Vo - 3 - 2 * 10-3 + = 0 1k + 2k 6k which yields Vo = 6 V. In addition, loop analysis applied as shown in Fig. 5.3d produces the equations I1 = -2 * 10-3 and 3k(I1 + I2) - 3 + 6kI2 = 0 which yield I2 = 1 mA and hence Vo = 6 V. EXAMPLE Consider now the network in Fig. 5.4a. Let us use superposition to find Vo˚. 5.4 + 6V ± – 6V 4 k⍀ 2 k⍀ ± – 4 k⍀ 2 k⍀ Vo 6 k⍀ 2 mA + V'o 6 k⍀ 2 k⍀ 2 k⍀ – – (a) (b) + 6 k⍀ + 6V ± – V1 + V'o 4 k⍀ 2 k⍀ – 4 k⍀ V"o 6 k⍀ – 2 k⍀ 2 mA 2 k⍀ – 2 k⍀ (c) (d) + 4 — k⍀ 3 + 6V 6 k⍀ 2 mA V"o Vo-6 2 k⍀ (e) Figure 5.4 I1 4 k⍀ V1 2 k⍀ I2 – Circuits used in Example 5.4. ± – I3 6 k⍀ Vo 2 k⍀ 2 mA – (f) irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 195 SECTION 5.2 The contribution of the 6-V source to V0 is found from the network in Fig. 5.4b, which is redrawn in Fig. 5.4c. The 2 k+6 k=8-k resistor and 4-k resistor are in parallel, and their combination is an 8/3-k resistor. Then, using voltage division, V1 = 6 ± 8 k 3 8 k + 2k 3 ≤ = SUPERPOSITION 195 SOLUTION 24 V 7 Applying voltage division again, Voœ = V1 a 6k 18 b = V 6k + 2k 7 The contribution of the 2-mA source is found from Fig. 5.4d, which is redrawn in Fig. 5.4e. Vofl is simply equal to the product of the current source and the parallel combination of the resistors; that is, Vofl = A2 * 10-3 B a 10 30 k兾兾6k b = V 3 7 Then 48 V 7 Vo = Voœ + Vofl = A nodal analysis of the network can be performed using Fig. 5.4f. The equation for the supernode is -2 * 10-3 + AVo - 6B - V1 2k + Vo - V1 Vo + = 0 4k 6k The equation for the node labeled V1 is V1 - AVo - 6B V1 - Vo V1 + + = 0 4k 2k 2k Solving these two equations, which already contain the constraint equation for the supernode, yields Vo = 48兾7 V. Once again, referring to the network in Fig. 5.4f, the mesh equations for the network are -6 + 4kAI1 - I3 B + 2kAI1 - I2 B = 0 I2 = 2 * 10-3 2kAI3 - I2 B + 4kAI3 - I1 B + 6kI3 = 0 Solving these equations, we obtain I3 = 8兾7 mA and, hence, Vo = 48兾7 V. Let us demonstrate the power of superposition in the analysis of op-amp circuits by determining the input/output relationship for the op-amp configuration shown in Fig. 5.5a. EXAMPLE The contribution of V1 to the output Vo is derived from the network in Fig. 5.5b where V2 is SOLUTION set to zero. This circuit is the basic inverting gain configuration and Vo1 R2 = V1 R1 The contribution due to V2 is shown in Fig. 5.5c where V1 is set to zero. This circuit is the basic noninverting configuration and Vo2 R2 = 1 + V2 R1 5.5 irwin05-189-244hr2.qxd 196 22-07-2010 CHAPTER 5 9:47 Page 196 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Therefore, using superposition, Vo = c 1 + R2 R2 d V2 - c d V1 R1 R1 Thus, in this case, we have used what we learned in Chapter 4, via superposition, to immediately derive the input/output relationship for the network in Fig. 5.5a. R2 Figure 5.5 (a) A superposition example circuit; (b) the circuit with V2 set to zero; (c) the circuit with V1 set to zero. R1 + + V1 ± – Vo1 R2 – R1 + V1 ± – (b) + R2 Vo V2 ± – – (a) + V2 ± – R1 R2 + Vo2 – (c) Problem-Solving Strategy Applying Superposition Step 1. In a network containing multiple independent sources, each source can be applied independently with the remaining sources turned off. Step 2. To turn off a voltage source, replace it with a short circuit, and to turn off a current source, replace it with an open circuit. Step 3. When the individual sources are applied to the circuit, all the circuit laws and techniques we have learned, or will soon learn, can be applied to obtain a solution. Step 4. The results obtained by applying each source independently are then added together algebraically to obtain a solution. Superposition can be applied to a circuit with any number of dependent and independent sources. In fact, superposition can be applied to such a network in a variety of ways. For example, a circuit with three independent sources can be solved using each source acting alone, as we have just demonstrated, or we could use two at a time and sum the result with irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 197 SECTION 5.2 SUPERPOSITION 197 that obtained from the third acting alone. In addition, the independent sources do not have to assume their actual value or zero. However, it is mandatory that the sum of the different values chosen add to the total value of the source. Superposition is a fundamental property of linear equations and, therefore, can be applied to any effect that is linearly related to its cause. In this regard it is important to point out that although superposition applies to the current and voltage in a linear circuit, it cannot be used to determine power because power is a nonlinear function. Learning Assessment E5.3 Compute Vo in circuit in Fig. E5.3 using superposition. 3 k⍀ 12 V ANSWER: Vo = + 4 k⍀ ± – 4 V. 3 2 mA 2 k⍀ Vo – Figure E5.3 ANSWER: V0 5.6 V. E5.4 Find V0 in Fig. E5.4 using superposition. 2 mA 12 V + 2 k⍀ – + 3 k⍀ 8 mA 6 k⍀ 1 k⍀ – Figure E5.4 ANSWER: I0 2/3 mA. E5.5 Find I0 in Fig. E5.5 using superposition. 6 mA 6 k⍀ 2 k⍀ 12 V + 4 k⍀ – 2 k⍀ 2 mA Figure E5.5 Vo Io 4 k⍀ irwin05-189-244hr2.qxd 198 22-07-2010 CHAPTER 5 5.3 9:47 Page 198 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Thus far we have presented a number of techniques for circuit analysis. At this point we will add two theorems to our collection of tools that will prove to be extremely useful. The theorems are named after their authors, M. L. Thévenin, a French engineer, and E. L. Norton, a scientist formerly with Bell Telephone Laboratories. Suppose that we are given a circuit and that we wish to find the current, voltage, or power that is delivered to some resistor of the network, which we will call the load. Thévenin’s theorem tells us that we can replace the entire network, exclusive of the load, by an equivalent circuit that contains only an independent voltage source in series with a resistor in such a way that the current–voltage relationship at the load is unchanged. Norton’s theorem is identical to the preceding statement except that the equivalent circuit is an independent current source in parallel with a resistor. Note that this is a very important result. It tells us that if we examine any network from a pair of terminals, we know that with respect to those terminals, the entire network is equivalent to a simple circuit consisting of an independent voltage source in series with a resistor or an independent current source in parallel with a resistor. In developing the theorems, we will assume that the circuit shown in Fig. 5.6a can be split into two parts, as shown in Fig. 5.6b. In general, circuit B is the load and may be linear or nonlinear. Circuit A is the balance of the original network exclusive of the load and must be linear. As such, circuit A may contain independent sources, dependent sources and resistors, or any other linear element. We require, however, that a dependent source and its control variable appear in the same circuit. Circuit A delivers a current i to circuit B and produces a voltage vo across the input terminals of circuit B. From the standpoint of the terminal relations of circuit A, we can replace circuit B by a voltage source of vo volts (with the proper polarity), as shown in Fig. 5.6c. Since the terminal voltage is unchanged and circuit A is unchanged, the terminal current i is unchanged. Now, applying the principle of superposition to the network shown in Fig. 5.6c, the total current i shown in the figure is the sum of the currents caused by all the sources in circuit A and the source vo that we have just added. Therefore, via superposition the current i can be written Thévenin’s and Norton’s Theorems i = io + isc 5.1 where io is the current due to vo with all independent sources in circuit A made zero (i.e., voltage sources replaced by short circuits and current sources replaced by open circuits), and isc is the short-circuit current due to all sources in circuit A with vo replaced by a short circuit. The terms io and vo are related by the equation io = -vo RTh 5.2 where RTh is the equivalent resistance looking back into circuit A from terminals A-B with all independent sources in circuit A made zero. Figure 5.6 Concepts used to develop Thévenin’s theorem. i Original circuit Circuit A A i + vo (linear) Circuit B A Circuit ± vo – A (linear) – (a) B B (b) (c) irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 199 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 199 Substituting Eq. (5.2) into Eq. (5.1) yields vo + isc RTh i = - 5.3 This is a general relationship and, therefore, must hold for any specific condition at terminals A-B. As a specific case, suppose that the terminals are open-circuited. For this condition, i=0 and vo is equal to the open-circuit voltage voc˚. Thus, Eq. (5.3) becomes i = 0 = -voc + isc RTh 5.4 Hence, voc = RTh˚isc 5.5 This equation states that the open-circuit voltage is equal to the short-circuit current times the equivalent resistance looking back into circuit A with all independent sources made zero. We refer to RTh as the Thévenin equivalent resistance. Substituting Eq. (5.5) into Eq. (5.3) yields i = -vo voc + RTh RTh or vo = voc - RTh˚i 5.6 Let us now examine the circuits that are described by these equations. The circuit represented by Eq. (5.6) is shown in Fig. 5.7a. The fact that this circuit is equivalent at terminals A-B to circuit A in Fig. 5.6 is a statement of Thévenin’s theorem. The circuit represented by Eq. (5.3) is shown in Fig. 5.7b. The fact that this circuit is equivalent at terminals A-B to circuit A in Fig. 5.6 is a statement of Norton’s theorem. Having demonstrated that there is an inherent relationship between the Thévenin equivalent circuit and the Norton equivalent circuit, we now proceed to apply these two important and useful theorems. The manner in which these theorems are applied depends on the structure of the original network under investigation. For example, if only independent sources are present, we can calculate the open-circuit voltage or short-circuit current and the Thévenin equivalent resistance. However, if dependent sources are also present, the Thévenin equivalent will be determined by calculating voc and isc˚, since this is normally the best approach for determining RTh in a network containing dependent sources. Finally, if circuit A contains no independent sources, then both voc and isc will necessarily be zero. (Why?) Thus, we cannot determine RTh by voc兾isc˚, since the ratio is indeterminate. We must look for another approach. Notice that if voc = 0, then the equivalent circuit is merely the unknown resistance RTh˚. If we apply an external source to circuit A—a test source vt—and determine the current, it˚, which i oc ± – RTh A i A + + vo (a) Circuit B isc RTh vo – – B B (b) Figure 5.7 Circuit B (a) Thévenin and (b) Norton equivalent circuits. irwin05-189-244hr2.qxd 200 22-07-2010 CHAPTER 5 9:47 Page 200 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S flows into circuit A from vt˚, then RTh can be determined from RTh = vt兾it˚. Although the numerical value of vt need not be specified, we could let vt = 1 V and then RTh = 1兾it˚. Alternatively, we could use a current source as a test source and let it = 1 A; then vt = (1)RTh˚. Before we begin our analysis of several examples that will demonstrate the utility of these theorems, remember that these theorems, in addition to being another approach, often permit us to solve several small problems rather than one large one. They allow us to replace a network, no matter how large, at a pair of terminals with a Thévenin or Norton equivalent circuit. In fact, we could represent the entire U.S. power grid at a pair of terminals with one of the equivalent circuits. Once this is done, we can quickly analyze the effect of different loads on a network. Thus, these theorems provide us with additional insight into the operation of a specific network. C I R C U I T S CO N TA I N I N G O N LY I N D E P E N D E N T S O U R C E S Let us use Thévenin’s and Norton’s theorems to find Vo in the network in Example 5.3. EXAMPLE 5.6 SOLUTION Figure 5.8 Circuits used in Example 5.6. 2 k⍀ The circuit is redrawn in Fig. 5.8a. To determine the Thévenin equivalent, we break the network at the 6-k load as shown in Fig. 5.8b. KVL indicates that the open-circuit voltage, Voc˚, is equal to 3 V plus the voltage V1˚, which is the voltage across the current source. The 2 mA from the current source flows through the two resistors (where else could it possibly go!) and, therefore, V1 = A2 * 10-3 B(1k + 2k) = 6 V. Therefore, Voc = 9 V. By making both sources zero, we can find the Thévenin equivalent resistance, RTh˚, using the circuit in Fig. 5.8c. Obviously, RTh = 3 k. Now our Thévenin equivalent circuit, consisting of Voc and RTh˚, is connected back to the original terminals of the load, as shown in Fig. 5.8d. Using a simple voltage divider, we find that Vo = 6 V. To determine the Norton equivalent circuit at the terminals of the load, we must find the short-circuit current as shown in Fig. 5.8e. Note that the short circuit causes the 3-V source to be directly across (i.e., in parallel with) the resistors and the current source. Therefore, I1 = 3兾(1k + 2k) = 1 mA. Then, using KCL, Isc = 3 mA. We have already determined RTh, and, therefore, connecting the Norton equivalent to the load results in the circuit in Fig. 5.8f. Hence, Vo is equal to the source current multiplied by the parallel resistor combination, which is 6 V. 2 k⍀ –± + 3V 1 k⍀ + Vo 6 k⍀ 2 mA 1 k⍀ 3 k⍀ 2 k⍀ – (d) 2 mA RTh 1 k⍀ – (c) + 3V Vo 6 k⍀ Voc –± + ± – + (b) (a) 9V 3V V1 – – 2 k⍀ –± 3 k⍀ 1 k⍀ I1 6 k⍀ Isc 3 mA 2 mA (e) Vo – (f) irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 201 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 201 Consider for a moment some salient features of this example. Note that in applying the theorems there is no point in breaking the network to the left of the 3-V source, since the resistors in parallel with the current source are already a Norton equivalent. Furthermore, once the network has been simplified using a Thévenin or Norton equivalent, we simply have a new network with which we can apply the theorems again. The following example illustrates this approach. EXAMPLE Let us use Thévenin’s theorem to find Vo in the network in Fig. 5.9a. Voc1 = 12 a 5.7 SOLUTION If we break the network to the left of the current source, the open-circuit voltage Voc1 is as shown in Fig. 5.9b. Since there is no current in the 2-k resistor and therefore no voltage across it, Voc1 is equal to the voltage across the 6-k resistor, which can be determined by voltage division as 6k b = 8V 6k + 3k The Thévenin equivalent resistance, RTh1˚, is found from Fig. 5.9c as RTh1 = 2k + (3k)(6k) 3k + 6k = 4 k Connecting this Thévenin equivalent back to the original network produces the circuit shown in Fig. 5.9d. We can now apply Thévenin’s theorem again, and this time we break the network to the right of the current source as shown in Fig. 5.9e. In this case Voc2 is Figure 5.9 Voc2 = A2 * 10-3 B(4k) + 8 = 16 V 3 k 2 k Circuits used in Example 5.7. 3 k 4 k 2 k + 12 V ± – 6 k Vo 8 k 2 mA + 12 V ± – – – (b) (a) 4 k 2 k 3 k Voc1 6 k 4 k + RTh1 6 k 8V ± – Vo 8 k 2 mA – (d) (c) 4 k 4 k 4 k 4 k + 8V ± – + Voc2 2 mA (e) RTh2 16 V ± – 8 k – Vo – (f) (g) irwin05-189-244hr2.qxd 202 22-07-2010 CHAPTER 5 9:47 Page 202 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S and RTh2 obtained from Fig. 5.9f is 4 k. Connecting this Thévenin equivalent to the remainder of the network produces the circuit shown in Fig. 5.9g. Simple voltage division applied to this final network yields Vo = 8 V. Norton’s theorem can be applied in a similar manner to solve this network; however, we save that solution as an exercise. EXAMPLE 5.8 SOLUTION It is instructive to examine the use of Thévenin’s and Norton’s theorems in the solution of the network in Fig. 5.4a, which is redrawn in Fig. 5.10a. If we break the network at the 6-k load, the open-circuit voltage is found from Fig. 5.10b. The equations for the mesh currents are -6 + 4kI1 + 2kAI1 - I2 B = 0 and I2 = 2 * 10-3 from which we easily obtain I1 = 5兾3 mA. Then, using KVL, Voc is Voc = 4kI1 + 2kI2 = 4k a = 5 * 10-3 b + 2kA2 * 10-3 B 3 32 V 3 RTh is derived from Fig. 5.10c and is R Th = (2k兾兾4k) + 2k = 10 k 3 Attaching the Thévenin equivalent to the load produces the network in Fig. 5.10d. Then using voltage division, we obtain Vo = = 32 3 ° 6k 6k + 10 ¢ k 3 48 V 7 In applying Norton’s theorem to this problem, we must find the short-circuit current shown in Fig. 5.10e. At this point the quick-thinking reader stops immediately! Three mesh equations applied to the original circuit will immediately lead to the solution, but the three mesh equations in the circuit in Fig. 5.10e will provide only part of the answer, specifically the short-circuit current. Sometimes the use of the theorems is more complicated than a straightforward attack using node or loop analysis. This would appear to be one of those situations. Interestingly, it is not. We can find Isc from the network in Fig. 5.10e without using the mesh equations. The technique is simple, but a little tricky, and so we ignore it at this time. Having said all these things, let us now finish what we have started. The mesh equations for the network in Fig. 5.10e are -6 + 4k˚AI1 - Isc B + 2k˚AI1 - 2 * 10-3 B = 0 2k˚AIsc - 2 * 10-3 B + 4k˚AIsc - I1 B = 0 irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 203 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 203 where we have incorporated the fact that I2 = 2 * 10-3 A. Solving these equations yields Isc = 16兾5 mA. RTh has already been determined in the Thévenin analysis. Connecting the Norton equivalent to the load results in the circuit in Fig. 5.10f. Solving this circuit yields Vo = 48兾7 V. + 6V ± – 4 k + 2 k 6 k I1 ± – 6V 4 k 2 k Vo Voc I2 2 k 2 mA – 2 k 2 mA – (b) (a) 4 k 2 k 10 — k 3 RTh + ± – 2 k 32 —V 3 6 k Vo – (c) 6V ± – (d) I1 4 k 2 k Isc I2 2 k + 16 — mA 5 10 — k 3 2 mA Figure 5.10 Vo – (e) Circuits used in Example 5.8. 6 k (f) irwin05-189-244hr4.qxd 204 12-08-2010 CHAPTER 5 14:56 Page 204 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Learning Assessments E5.6 Use Thévenin’s theorem to find Vo in the network in Fig. E5.6. ANSWER: Vo 3V. 2 k 6 k + 3 k 6V ± – 4 k + Vo 12 V – Figure E5.6 4 V. 3 E5.7 Find Vo in the circuit in Fig. E5.6 using both Thévenin’s and Norton’s theorems. When ANSWER: Vo = E5.8 Find Vo in Fig. E5.8 using Thévenin’s theorem. ANSWER: V0 3.88 V. deriving the Norton equivalent circuit, break the network to the left of the 4-k resistor. Why? 2 mA 6 k 2 k + 6 k 8 mA 1 k 3 k + – 12 V – Figure E5.8 ANSWER: I0 0.857 mA. E5.9 Find I0 in Fig. E5.9 using Norton’s theorem. 3 k 4 k 2 mA - 12 V + 2 k 6 k 3 k 8V + + 10 V Figure E5.9 Vo I0 6 k irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 205 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 205 C I R C U I T S CO N TA I N I N G O N LY D E P E N D E N T S O U R C E S As we have stated earlier, the Thévenin or Norton equivalent of a network containing only dependent sources is RTh˚. The following examples will serve to illustrate how to determine this Thévenin equivalent resistance. We wish to determine the Thévenin equivalent of the network in Fig. 5.11a at the terminals A-B. – Vx + – 1 k 2 k ± – 5.9 + 1 k 1 k A ± 1 k Vx EXAMPLE 2Vx V1 2 k 2 k 1 k I2 I1 A Io 1 k ± – 2Vx 1V I3 – B (a) ± – 2 k B (b) Figure 5.11 Networks employed in Example 5.9. Our approach to this problem will be to apply a 1-V source at the terminals as shown in Fig. 5.11b and then compute the current Io and RTh 1/Io. The equations for the network in Fig. 5.11b are as follows. KVL around the outer loop specifies that V1 Vx 1 The KCL equation at the node labeled V1 is V1 - 2Vx V1 V1 - 1 + + = 0 1k 2k 1k Solving the equations for Vx yields Vx = 3兾7 V. Knowing Vx˚, we can compute the currents I1˚, I2˚, and I3˚. Their values are Vx 3 = mA 1k 7 1 - 2Vx 1 = mA I2 = 1k 7 I1 = I3 = 1 1 = mA 2k 2 Therefore, Io = I1 + I2 + I3 = 15 mA 14 and 1 Io 14 = k 15 RTh = SOLUTION irwin05-189-244hr2.qxd 206 22-07-2010 CHAPTER 5 9:47 Page 206 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S EXAMPLE Let us determine RTh at the terminals A-B for the network in Fig. 5.12a. 5.10 SOLUTION Our approach to this problem will be to apply a 1-mA current source at the terminals A-B and compute the terminal voltage V2 as shown in Fig. 5.12b. Then RTh = V2兾0.001. The node equations for the network are V1 - 2000Ix V1 V1 - V2 + + = 0 2k 1k 3k V2 - V1 V2 + = 1 * 10-3 3k 2k and Ix = V1 1k Solving these equations yields 10 V 7 V2 = and hence, RTh = Figure 5.12 Networks used in Example 5.10. = 2 k 3 k V2 1 * 10-3 10 k 7 V1 A V2 2 k ± – 2000Ix 1 k ± – 2 k A 3 k 2000Ix 1 k Ix 2 k 1 mA Ix B B (a) (b) C I R C U I T S CO N TA I N I N G B OT H I N D E P E N D E N T A N D D E P E N D E N T S O U R C E S In these types of circuits we must calculate both the open-circuit voltage and short-circuit current to calculate the Thévenin equivalent resistance. Furthermore, we must remember that we cannot split the dependent source and its controlling variable when we break the network to find the Thévenin or Norton equivalent. We now illustrate this technique with a circuit containing a current-controlled voltage source. EXAMPLE 5.11 SOLUTION Let us use Thévenin’s theorem to find Vo in the network in Fig. 5.13a. To begin, we break the network at points A-B. Could we break it just to the right of the 12-V source? No! Why? The open-circuit voltage is calculated from the network in Fig. 5.13b. Note that we now use the source 2000Ixœ because this circuit is different from that in Fig. 5.13a. KCL for the supernode around the 12-V source is AVoc + 12B - A-2000I xœ B 1k + Voc + 12 Voc + = 0 2k 2k irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 207 SECTION 5.3 207 THÉVENIN’S AND NORTON’S THEOREMS where I xœ = Voc 2k yielding Voc = -6 V . Isc can be calculated from the circuit in Fig. 5.13c. Note that the presence of the short circuit forces I xfl to zero and, therefore, the network is reduced to that shown in Fig. 5.13d. Therefore, Isc = -12 = -18 mA 2 k 3 Then RTh = Voc 1 = k Isc 3 Connecting the Thévenin equivalent circuit to the remainder of the network at terminals A-B produces the circuit in Fig. 5.13e. At this point, simple voltage division yields Vo = (-6) 12 V + 2000Ix 1k 1 ¢ 1k + 1k + k 3 = -18 V 7 Figure 5.13 Circuits used in Example 5.11. 12 V A ±– 1 k ° + 1 k 2 k 2 k 1 k Ix Vo 1 k + 2000I'x 2 k I'x – (b) 12 V 12 V A ±– 2000I"x 2 k ±– Isc 2 k 1 k A Isc 2 k I"x B (c) + + 1 k 1 k 6V Vo – B (e) B (d) A 1 — k 3 Voc – B (a) + + 2 k B 1 k A ±– irwin05-189-244hr2.qxd 208 22-07-2010 CHAPTER 5 9:47 Page 208 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S EXAMPLE Let us find Vo in the network in Fig. 5.14a using Thévenin’s theorem. 5.12 SOLUTION Voc is determined from the network in Fig. 5.14b. Note that I1 = Vxœ 2k I2 = 2 mA and Vxœ = 4k a Vxœ - 2 * 10-3 b 2k Solving these equations yields I1 = 4 mA and, hence, Voc = 2kI1 + 3 = 11 V Isc is derived from the circuit in Fig. 5.14c. Note that if we collapse the short circuit, the network is reduced to that in Fig. 5.14d. Although we have temporarily lost sight of Isc˚, we can + Vx — 2000 + V'x — 2000 2 k – Vx + I1 – Vo 6 k V'x + Voc 4 k 4 k ± – 2 k 3V ± – I2 2 mA 3V 2 mA – – (a) V"x — (b) 2 k 2000 – V"x + I3 Isc 4 k 2 mA ± – V"x — 2 mA 3V – V"x 4 k (c) (d) + 2 k + - 11 V 6 k Figure 5.14 Circuits used in Example 5.12. Vo – (e) I3 2000 2 k + + 3V irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 209 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 209 easily find the branch currents and they, in turn, will yield Isc˚. KCL at the node at the bottom left of the network is V xfl V xfl = - 2 * 10-3 4k 2000 or V xfl = 8 V Then since I3 = 3 3 = mA 2k 2 as shown in Fig. 5.14c, Isc = = Then RTh = Vxfl + I3 2000 11 mA 2 Voc = 2 k Isc Connecting the Thévenin equivalent circuit to the remainder of the original network produces the circuit in Fig. 5.14e. Simple voltage division yields Vo = 11 a = 6k b 2k + 6k 33 V 4 We will now reexamine a problem that was solved earlier using both nodal and loop analyses. The circuit used in Examples 3.10 and 3.20 is redrawn in Fig. 5.15a. Since a dependent source is present, we will have to find the open-circuit voltage and the short-circuit current in order to employ Thévenin’s theorem to determine the output voltage Vo˚. As we begin the analysis, we note that the circuit can be somewhat simplified by first forming a Thévenin equivalent for the leftmost and rightmost branches. Note that these two branches are in parallel and neither branch contains the control variable. Thus, we can simplify the network by reducing these two branches to one via a Thévenin equivalent. For the circuit shown in Fig. 5.15b, the open-circuit voltage is Voc1 = 2 (1k) + 4 = 6 V k And the Thévenin equivalent resistance at the terminals, obtained by looking into the terminals with the sources made zero, is R Th1 = 1 k The resultant Thévenin equivalent circuit is now connected to the remaining portion of the circuit producing the network in Fig. 5.15c. Now we break the network shown in Fig. 5.15c at the output terminals to determine the open-circuit voltage Voc2 as shown in Fig. 5.15d. Because of the presence of the voltage sources, we will use a nodal analysis to find the open-circuit voltage with the help of a supernode. The node equations for this network are V1 = 3Vxœ V1 - 2Vxœ V1 - 6 2 + = 1k 1k k EXAMPLE 5.13 SOLUTION irwin05-189-244hr2.qxd 210 22-07-2010 CHAPTER 5 9:47 Page 210 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S + ± – 2Vx Vo 1 k + 1 k – 2 —A k 1 k 2 —A k 1 k + Vx 2 —A k 1 k ± – Voc1 ± – 4V 4V – – (a) (b) V1 + + 6V ± – ± – 2Vx Vo 1 k ± – 6V ± – 2V'x Voc2 – + Vx 1 k – 1 k + 2 —A k 1 k V'x 1 k – ± – 2V"x 6V 2 —A k Isc2 6V 2V V"x 2 —A k 1 k 2 —A k – (e) (f) 1 k + 2V ± – 1 k Vo – (g) Figure 5.15 Circuits used in Example 5.13. 2 A Isc2=— k 4 —A k 1 k + 1 k 2 —A k (d) 0A ± – 1 k – (c) 6V 1 k 6V 2 —A k irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 211 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 211 and thus Vxœ = 2 V and V1 = 6 V. Then, the open-circuit voltage, obtained using the KVL equation -2V xœ + Voc2 + 2 (1k) = 0 k is Voc2 = 2 V The short-circuit current is derived from the network shown in Fig. 5.15e. Once again we employ the supernode, and the network equations are V2 = 3Vxfl V2 - 2Vxfl V2 - 6 2 + = 1k 1k k The node voltages obtained from these equations are Vxfl = 2 V and V2 = 6 V. The line diagram shown in Fig. 5.15f displays the node voltages and the resultant branch currents. (Node voltages are shown in the circles, and branch currents are identified with arrows.) The node voltages and resistors are used to compute the resistor currents, while the remaining currents are derived by KCL. As indicated, the short-circuit current is Isc2 = 2 mA Then, the Thévenin equivalent resistance is R Th2 = Voc2 Isc2 = 1 k The Thévenin equivalent circuit now consists of a 2-V source in series with a 1-k resistor. Connecting this Thévenin equivalent circuit to the load resistor yields the network shown in Fig. 5.15g. A simple voltage divider indicates that Vo = 1 V. Problem-Solving Strategy Step 1. Remove the load and find the voltage across the open-circuit terminals, Voc˚. All the circuit analysis techniques presented here can be used to compute this voltage. Step 2. Determine the Thévenin equivalent resistance of the network at the open terminals with the load removed. Three different types of circuits may be encountered in determining the resistance, RTh˚. (a) If the circuit contains only independent sources, they are made zero by replacing the voltage sources with short circuits and the current sources with open circuits. RTh is then found by computing the resistance of the purely resistive network at the open terminals. (b) If the circuit contains only dependent sources, an independent voltage or current source is applied at the open terminals and the corresponding current or voltage at these terminals is measured. The voltage/current ratio at the terminals is the Thévenin equivalent resistance. Since there is no energy source, the open-circuit voltage is zero in this case. (c) If the circuit contains both independent and dependent sources, the open-circuit terminals are shorted and the short-circuit current between these terminals is determined. The ratio of the open-circuit voltage to the short-circuit current is the resistance RTh˚. Step 3. If the load is now connected to the Thévenin equivalent circuit, consisting of Voc in series with RTh˚, the desired solution can be obtained. The problem-solving strategy for Norton’s theorem is essentially the same as that for Thévenin’s theorem with the exception that we are dealing with the short-circuit current instead of the open-circuit voltage. Applying Thévenin’s Theorem irwin05-189-244hr2.qxd 212 22-07-2010 CHAPTER 5 15:11 Page 212 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Learning Assessment E5.10 Find Vo in the circuit in Fig. E5.10 using Thévenin’s theorem. + Vx ANSWER: Vo = 36 V. 13 – + 4 k⍀ 4 k⍀ 12 V ± – Vo 6 k⍀ x - V + — 2 – Figure E5.10 ANSWER: V0 ⫽ 6.29 V. E5.11 Find Vo in Fig. E5.11 using Thévenin’s theorem. 2 mA 4 kIx 2 k⍀ + - + 3 k⍀ 8 mA IX 6 k⍀ 1 k⍀ Vo – Figure E5.11 E5.12 Use Thévenin’s theorem to find the power supplied by the 12-V source in Fig. E5.12. ANSWER: 8.73 mW. 2 mA 6 k⍀ 2 k⍀ 6 k⍀ 3 k⍀ 8 mA 12 V 1 k⍀ + – Figure E5.12 E5.13 Find the Thévenin equivalent of the network at terminals A ⫺ B in Fig. E5.13. 3 k⍀ – Vx + A 1 k⍀ 2 k⍀ Vx — 2000 2 k⍀ Figure E5.13 B ANSWER: RTh ⫽ 1619 ⍀. irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 213 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 213 Having examined the use of Thévenin’s and Norton’s theorems in a variety of different types of circuits, it is instructive to look at yet one other aspect of these theorems that we find useful in circuit analysis and design. This additional aspect can be gleaned from the Thévenin equivalent and Norton equivalent circuits. The relationships specified in Fig. 5.7 and Eq. (5.5) have special significance because they represent what is called a source transformation or source exchange. What these relationships tell us is that if we have embedded within a network a current source i in parallel with a resistor R, we can replace this combination with a voltage source of value v=iR in series with the resistor R. The reverse is also true; that is, a voltage source v in series with a resistor R can be replaced with a current source of value i=v/R in parallel with the resistor R. Parameters within the circuit (e.g., an output voltage) are unchanged under these transformations. We must emphasize that the two equivalent circuits in Fig. 5.7 are equivalent only at the two external nodes. For example, if we disconnect circuit B from both networks in Fig. 5.7, the equivalent circuit in Fig. 5.7b dissipates power, but the one in Fig. 5.7a does not. EXAMPLE We will now demonstrate how to find Vo in the circuit in Fig. 5.16a using the repeated application of source transformation. 5.14 SOLUTION If we begin at the left end of the network in Fig. 5.16a, the series combination of the 12-V source and 3-k resistor is converted to a 4-mA current source in parallel with the 3-k resistor. If we combine this 3-k resistor with the 6-k resistor, we obtain the circuit in Fig. 5.16b. Note that at this point we have eliminated one circuit element. Continuing the reduction, we convert the 4-mA source and 2-k resistor into an 8-V source in series with this same 2-k resistor. The two 2-k resistors that are in series are now combined to produce the network in Fig. 5.16c. If we now convert the combination of the 8-V source and 4-k resistor into a 2-mA source in parallel with the 4-k resistor and combine the resulting current source with the other 2-mA source, we arrive at the circuit shown in Fig. 5.16d. At this point, we can simply apply current division to the two parallel resistance paths and obtain Io = A4 * 10-3 B a and hence, 4k b = 1 mA 4k + 4k + 8k Vo = A1 * 10-3 B(8k) = 8 V Figure 5.16 The reader is encouraged to consider the ramifications of working this problem using any of the other techniques we have presented. 2 k 3 k 12 V ± – + 4 k 6 k 2 mA 8 k Vo Circuits used in Example 5.14. 4 mA 2 k 8 k 2 mA – (b) Io 8V ± – + 4 k 2 mA 8 k Vo + 4 k 4 mA 4 k – (c) Vo – (a) 4 k + 4 k 2 k 8 k Vo – (d) irwin05-189-244hr2.qxd 214 22-07-2010 CHAPTER 5 9:47 Page 214 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Note that this systematic, sometimes tedious, transformation allows us to reduce the network methodically to a simpler equivalent form with respect to some other circuit element. However, we should also realize that this technique is worthless for circuits of the form shown in Fig. 5.4. Furthermore, although applicable to networks containing dependent sources, it is not as useful as other techniques, and care must be taken not to transform the part of the circuit that contains the control variable. Learning Assessment 4 V. 3 E5.14 Find Vo in the circuit in Fig. E5.3 using source exchange. ANSWER: Vo = E5.15 Find the Io in Fig. E5.15 using source transformations. ANSWER: Io 1.94 mA. 6 k 2 k Io 6 k 8 mA 3 k 1 k 12 V 4 mA – + Figure E5.15 At this point let us pause for a moment and reflect on what we have learned; that is, let us compare the use of node or loop analysis with that of the theorems discussed in this chapter. When we examine a network for analysis, one of the first things we should do is count the number of nodes and loops. Next we consider the number of sources. For example, are there a number of voltage sources or current sources present in the network? All these data, together with the information that we expect to glean from the network, give a basis for selecting the simplest approach. With the current level of computational power available to us, we can solve the node or loop equations that define the network in a flash. With regard to the theorems, we have found that in some cases the theorems do not necessarily simplify the problem and a straightforward attack using node or loop analysis is as good an approach as any. This is a valid point provided that we are simply looking for some particular voltage or current. However, the real value of the theorems is the insight and understanding that they provide about the physical nature of the network. For example, superposition tells us what each source contributes to the quantity under investigation. However, a computer solution of the node or loop equations does not tell us the effect of changing certain parameter values in the circuit. It does not help us understand the concept of loading a network or the ramifications of interconnecting networks or the idea of matching a network for maximum power transfer. The theorems help us to understand the effect of using a transducer at the input of an amplifier with a given input resistance. They help us explain the effect of a load, such as a speaker, at the output of an amplifier. We derive none of this information from a node or loop analysis. In fact, as a simple example, suppose that a network at a specific pair of terminals has a Thévenin equivalent circuit consisting of a voltage source in series with a 2-k resistor. If we connect a 2- resistor to the network at these terminals, the voltage across the 2- resistor will be essentially nothing. This result is fairly obvious using the Thévenin theorem approach; however, a node or loop analysis gives us no clue as to why we have obtained this result. We have studied networks containing only dependent sources. This is a very important topic because all electronic devices, such as transistors, are modeled in this fashion. Motors in power systems are also modeled in this way. We use these amplification devices for many different purposes, such as speed control for automobiles. In addition, it is interesting to note that when we employ source transformation as we did in Example 5.14, we are simply converting back and forth between a Thévenin equivalent circuit and a Norton equivalent circuit. irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 215 SECTION 5.3 THÉVENIN’S AND NORTON’S THEOREMS 215 Finally, we have a powerful tool at our disposal that can be used to provide additional insight and understanding for both circuit analysis and design. That tool is Microsoft Excel, and it permits us to study the effects, on a network, of varying specific parameters. The following example will illustrate the simplicity of this approach. We wish to use Microsoft Excel to plot the Thévenin equivalent parameters Voc and RTh for the circuit in Fig. 5.17 over the Rx range 0 to 10 k. 6V 12 V ±– + + 5.15 Figure 5.17 Circuit used in Example 5.15. Voc Rx 4 k RTh EXAMPLE - The Thévenin resistance is easily found by replacing the voltage sources with short circuits. The result is RTh = 4兾兾Rx = 4Rx 4 + Rx 5.7 where Rx and RTh are in k. Superposition can be used effectively to find Voc˚. If the 12-V source is replaced by a short circuit Voc1 = -6 c Rx d Rx + 4 Applying this same procedure for the 6-V source yields Voc2 = 12 and the total open-circuit voltage is Voc = 12 - 6 c Rx d Rx + 4 5.8 In Excel we wish to (1) vary Rx between 0 and 10 k, (2) calculate RTh and Voc at each Rx value, and (3) plot Voc and RTh versus Rx˚. We begin by opening Excel and entering column headings as shown in Fig. 5.18a. Next, we enter a zero in the first cell of the Rx column at column-row location A4. To automatically fill the column with values, go to the Edit menu and select Fill/Series to open the window shown in Fig. 5.18b, which has already been edited appropriately for 101 data points. The result is a series of Rx values from 0 to 10 k in 100 steps. To enter Eq. (5.8), go to location B4 (right under the Voc heading). Enter the following text and do not forget the equal sign: =12-6*A4/(A4+4) This is Eq. (5.8) with Rx replaced by the first value for Rx˚, which is at column-row location A4. Similarly for RTh˚, enter the following expression at C4. =4*A4/(A4+4) To replicate the expression in cell B4 for all Rx values, select cell B4, grab the lower right corner of the cell, hold and drag down to cell B104, and release. Repeat for RTh by replicating cell C4. To plot the data, first drag the cursor across all cells between A4 and C104. Next, from the Insert menu, select Chart. We recommend strongly that you choose the XY (Scatter) SOLUTION irwin05-189-244hr2.qxd 216 22-07-2010 CHAPTER 5 9:47 Page 216 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S chart type. Excel will take you step by step through the basic formatting of your chart, which, after some manipulations, might look similar to the chart in Fig. 5.18c. (b) 15 3 (a) 2.5 10 2 7.5 1.5 5 2.5 0.5 0 Figure 5.18 0 0 2.5 (a) The Excel spreadsheet for Example 5.15 showing the desired column headings. (b) The Fill/Series window edited for varying Rx and (c) the final plot of Voc and RTh. 5.4 Maximum Power Transfer 7.5 5 10 R (kΩ) (c) In circuit analysis we are sometimes interested in determining the maximum power that can be delivered to a load. By employing Thévenin’s theorem, we can determine the maximum power that a circuit can supply and the manner in which to adjust the load to effect maximum power transfer. Suppose that we are given the circuit shown in Fig. 5.19. The power that is delivered to the load is given by the expression Pload = i2RL = a i 1 Voc RTh RTh (kΩ) Voc(V) 12.5 2 v b RL R + RL We want to determine the value of RL that maximizes this quantity. Hence, we differentiate this expression with respect to RL and equate the derivative to zero: R AR + RL B v˚2 - 2v˚2RL AR + RL B dPload = = 0 dRL AR + RL B 4 2 v ± – RL which yields RL = R Figure 5.19 Equivalent circuit for examining maximum power transfer. In other words, maximum power transfer takes place when the load resistance RL = R. Although this is a very important result, we have derived it using the simple network in Fig. 5.19. However, we should recall that v and R in Fig. 5.19 could represent the Thévenin equivalent circuit for any linear network irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 217 SECTION 5.4 MAXIMUM POWER TRANSFER Let us find the value of RL for maximum power transfer in the network in Fig. 5.20a and the maximum power that can be transferred to this load. To begin, we derive the Thévenin equivalent circuit for the network exclusive of the load. Voc can be calculated from the circuit in Fig. 5.20b. The mesh equations for the network are 217 EXAMPLE 5.16 SOLUTION I1 = 2 * 10-3 3kAI2 - I1 B + 6kI2 + 3 = 0 Solving these equations yields I2 = 1兾3 mA and, hence, Voc = 4kI1 + 6kI2 = 10 V RTh˚, shown in Fig. 5.20c, is 6 k; therefore, RL = RTh = 6 k for maximum power transfer. The maximum power transferred to the load in Fig. 5.20d is PL = a 10 2 25 mW b (6k) = 12k 6 Voc + Figure 5.20 – RL 4 k 6 k 6 k I1 I2 ± – 3 k 2 mA 4 k 3V 3 k 2 mA (a) Circuits used in Example 5.16. ± – 3V (b) RTh RL=6 k 4 k 6 k 6 k 3 k 10 V ±– (c) (d) Let us find RL for maximum power transfer and the maximum power transferred to this load in the circuit in Fig. 5.21a. We wish to reduce the network to the form shown in Fig. 5.19. We could form the Thévenin equivalent circuit by breaking the network at the load. However, close examination of the network indicates that our analysis will be simpler if we break the network to the left of the 4-k resistor. When we do this, however, we must realize that for maximum power transfer RL = RTh + 4 k. Voc can be calculated from the network in Fig. 5.21b. Forming a supernode around the dependent source and its connecting nodes, the KCL equation for this supernode is Voc - 2000Ixœ Voc + (-4 * 10-3) + = 0 1k + 3k 2k EXAMPLE 5.17 SOLUTION irwin05-189-244hr2.qxd 218 22-07-2010 CHAPTER 5 9:47 Page 218 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S where Ixœ = Voc 2k These equations yield Voc 8 V. The short-circuit current can be found from the network in Fig. 5.21c. It is here that we find the advantage of breaking the network to the left of the 4-k resistor. The short circuit shorts the 2-k resistor and, therefore, I xfl = 0. Hence, the circuit is reduced to that in Fig. 5.21d, where clearly Isc = 4 mA. Then RTh = Voc = 2 k Isc Connecting the Thévenin equivalent to the remainder of the original circuit produces the network in Fig. 5.21e. For maximum power transfer RL = RTh + 4 k = 6 k, and the maximum power transferred is PL = a Figure 5.21 Circuits used in Example 5.17. –+ 1 k 8 8 2 b (6k) = mW 12k 3 2000 Ix 3 k RL 2 k 4 mA –+ 1 k 4 k 2 k Voc I'x – 3 k 4 mA Ix (a) (b) –+ 1 k + 2000 I'x 1 k 2000 I"x 3 k Isc 2 k 4 mA 3 k 4 mA Isc I"x (d) (c) 2 k + – 4 k RL=6 k 8V (e) Learning Assessment E5.16 Given the circuit in Fig. E5.16, find RL for maximum power transfer and the maximum power transferred. 6V 2 k + 6 k Figure E5.16 12 k RL ANSWER: RL = 6 k; 2 PL = mW. 3 irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 219 SECTION 5.4 MAXIMUM POWER TRANSFER E5.17 Find RL for maximum transfer and the maximum power transferred to RL in Fig E5.17. 3 k 12 V 219 ANSWER: 14/9 k, 2/7 mW. 4 k – + 2 mA 2 k RL Figure E5.17 E5.18 Find RL for maximum transfer and the maximum power transferred to RL in Fig E5.18. + Vx ANSWER: 24/13 k, 27/26 mW. – 4 k 12 V + – 4 k 2 RL – + Vx 6 k Figure E5.18 Given the network in Fig. 5.22 with Vin = 5 V and R1 = 2 , let us graphically examine a variety of aspects of maximum power transfer by plotting the parameters Vout˚, I, Pout˚, Pin and the efficiency = Pout兾Pin as a function of the resistor ratio R2兾R1˚. The parameters to be plotted can be determined by simple circuit analysis techniques. By voltage division Vout = c EXAMPLE 5.18 SOLUTION R2 R2 dV = c d (5) R 1 + R 2 in 2 + R2 From Ohm’s law I = Vin 5 = R1 + R2 2 + R2 The input and output powers are Pin = IVin = V2in 25 = R1 + R2 2 + R2 ˚˚P out = IVout = R2 c 2 2 Vin 5 d = R2 c d R1 + R2 2 + R2 R1=2 Figure 5.22 I + Vin=5 V ± – R2 Vout – Circuit used in maximum power transfer analysis. irwin05-189-244hr2.qxd 220 22-07-2010 CHAPTER 5 9:47 Page 220 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Finally, the efficiency is efficiency = Pout R2 R2 = = Pin R1 + R2 2 + R2 The resulting plots of the various parameters are shown in Fig. 5.23 for R2 ranging from 0.1R1 to 10R1˚. Note that as R2 increases, Vout increases toward Vin (5 V) as dictated by voltage division. Also, the current decreases in accordance with Ohm’s law. Thus, for small values of R2˚, Vout is small, and when R2 is large, I is small. As a result, the output power (the product of these two parameters) has a maximum at R2兾R1 = 1 as predicted by maximum power transfer theory. Maximum power does not correspond to maximum output voltage, current, or efficiency. In fact, at maximum power transfer, the efficiency is always 0.5, or 50%. If you are an electric utility supplying energy to your customers, do you want to operate at maximum power transfer? The answer to this question is an obvious “No” because the efficiency is only 50%. The utility would only be able to charge its customers for one-half of the energy produced. It is not uncommon for a large electric utility to spend billions of dollars every year to produce electricity. The electric utility is more interested in operating at maximum efficiency. Figure 5.23 5 Vout Max. Power Transfer Parameters (V, A, W) Maximum power transfer parameter plots for the network in Fig. 5.22. (The units for voltage, current, and power have volts, amperes, and watts, respectively.) 4 Pout 3 2 Current Pout/Pin 1 0 0 2 4 6 8 10 R2/R1 5.5 Application Example APPLICATION EXAMPLE 5.19 On Monday afternoon, Connie suddenly remembers that she has a term paper due Tuesday morning. When she sits at her computer to start typing, she discovers that the computer mouse doesn’t work. After disassembly and some inspection, she finds that the mouse contains a printed circuit board that is powered by a 5-V supply contained inside the computer case. Furthermore, the board is found to contain several resistors, some op-amps, and one unidentifiable device, which is connected directly to the computer’s 5-V supply as shown in Fig. 5.24a. Using a voltmeter to measure the node voltages, Connie confirms that all resis- irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 221 SECTION 5.6 DESIGN EXAMPLES 221 tors and op-amps are functioning properly and the power supply voltage reaches the mouse board. However, without knowing the mystery device’s function within the circuit, she cannot determine its condition. A phone call to the manufacturer reveals that the device is indeed linear but is also proprietary. With some persuasion, the manufacturer’s representative agrees that if Connie can find the Thévenin equivalent circuit for the element at nodes A-B with the computer on, he will tell her if it is functioning properly. Armed with a single 1-k resistor and a voltmeter, Connie attacks the problem. To find the Thévenin equivalent for the unknown device, together with the 5-V source, Connie first isolates nodes A and B from the rest of the devices on the board to measure the open-circuit voltage. The resulting voltmeter reading is VAB 2.4 V. Thus, the Thévenin equivalent voltage is 2.4 V. Then she connects the 1-k resistor at nodes A-B as shown in Fig. 5.24b. The voltmeter reading is now VAB = 0.8 V. Using voltage division to express VAB in terms of VTh˚, RTh, and Rtest in Fig. 5.24b yields the expression 0.8 = VTh a Unknown element 1k b 1k + RTh RTh A V=5 V ± – SOLUTION VTh ± – RTh=2 k A Rtest=1 k VAB=0.8 V ± – VTh=2.4 V B B B (a) (b) Solving the equations for RTh˚, we obtain RTh = 2.0 k A (c) Figure 5.24 Network used in Example 5.19. Therefore, the unknown device and the 5-V source can be represented at the terminals A-B by the Thévenin equivalent circuit shown in Fig. 5.24c. When Connie phones the manufacturer with the data, the representative informs her that the device has indeed failed. 5.6 Design Examples We often find that in the use of electronic equipment, there is a need to adjust some quantity such as voltage, frequency, contrast, or the like. For very accurate adjustments, it is most convenient if coarse and fine-tuning can be separately adjusted. Therefore, let us design a circuit in which two inputs (i.e., coarse and fine voltages) are combined to produce a new voltage of the form DESIGN EXAMPLE 5.20 1 1 Vtune = c d Vcoarse + c d Vfine 2 20 Because the equation to be realized is the sum of two terms, the solution appears to be an excellent application for superposition. Since the gain factors in the equation (i.e., 1/2 and 1/20) SOLUTION irwin05-189-244hr2.qxd 222 22-07-2010 CHAPTER 5 9:47 Page 222 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S are both less than one, a voltage divider with two inputs would appear to be a logical choice. A typical circuit for this application is shown in Fig. 5.25a. The two superposition subcircuits are shown in Figs. 5.25b and c. Employing voltage division in the network in Fig. 5.25b yields Vtune_C Vcoarse = c and therefore, R兾兾R 2 AR兾兾R 2 B + R 1 d = 1 2 R兾兾R 2 = R 1 In a similar manner, we find that Vtune_F Vfine which requires that R1 + R R兾兾R 1 AR兾兾R 1 B + R 2 d = 1 20 R2 = 19AR兾兾R1 B R2 Vcoarse ± – = c R1 ± Vfine – Vtune R2 Vcoarse ± – + Vtune_C R + R2 R1 - (a) ± Vfine – R Vtune_F (c) (b) Figure 5.25 (a) The coarse/fine adjustment circuit, (b) with Vfine set to zero and (c) with Vcoarse set to zero. • DESIGN EXAMPLE 5.21 Note that the two constraint equations for the resistors have three unknowns—R, R1, and R2˚. Thus, we must choose one resistor value and then solve for the two remaining values. If we arbitrarily select R = 1 k, then R1 = 900 and R2 = 9 k. This completes the design of the circuit. This example indicates that superposition is not only a useful analysis tool but provides insight into the design of new circuits. Coaxial cable is often used in very-high-frequency systems. For example, it is commonly used for signal transmission with cable television. In these systems resistance matching, the kind we use for maximum power transfer, is critical. In the laboratory, a common apparatus used in high-frequency research and development is the attenuator pad. The attenuator pad is basically a voltage divider, but the equivalent resistance at both its input ports is carefully designed for resistance matching. Given the network in Fig. 5.26 in which a source, modeled by VS and RS (50 ), drives an attenuator pad, which is connected to an equivalent load. Attenuator pad Figure 5.26 RS The model circuit for the attenuator pad design. R2 R2 50 VS ± – RTh-in RL R1 + Vout 50 RTh-out irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 223 SECTION 5.6 DESIGN EXAMPLES 223 Let us design the pad so that it has an equivalent resistance of 50 and divides (i.e., attenuates) the input voltage by a factor of 10. Since the attenuator or “T-Network” must have an equivalent resistance of 50 , we require that RTh-in and RTh-out be 50 . Since these Thévenin resistance values are the same and the circuit is symmetric, we can use the label R2 twice to indicate that those resistors will be the same value. SOLUTION RTh-in = R2 + C R1兾兾AR2 + 50B D = 50 RTh-out = R2 + C R1兾兾AR2 + 50B D = 50 Since the equations are identical, we refer to both Thévenin equivalent resistance parameters simply as RTh˚. The Thévenin equivalent voltage, VTh˚, can be easily derived from the circuit in Fig. 5.27 a using voltage division. VTh = VS c RS R2 R1 d R 1 + R 2 + 50 R2 50 + VS ± – Figure 5.27 RTh VTh R1 + VTh ± – RL Vout 50 - (a) The circuit used in finding VTh and (b) the resulting model. - RTh (a) (b) From the Thévenin equivalent circuit in Fig. 5.27b, we find Vout = VTh c VTh 50 d = R Th + 50 2 Combining these equations yields the attenuation from VS to Vout Vout Vout VTh R1 1 1 = c dc d = c d = VS VTh VS 2 R 1 + R 2 + 50 10 The Thévenin equivalent resistance equation and this attenuation equation provide us with two equations in the two unknowns R1 and R2˚. Solving these equations yields R1 = 20.83 and R2 = 33.33 . For precise resistance matching, these resistors must be very accurate. With such low resistor values, the power dissipation can become significant as VS is increased. For example, if VS = 10 V, Vout = 1 V and the power dissipated in the R2 resistor connected to the input source is 333 mW. To keep the temperature of that resistor at reasonable levels, the power rating of that resistor should be at least 0.5 W. Let us design a circuit that will realize the following equation: Vo = -3VS - 2000IS An examination of this equation indicates that we need to add two terms, one of which is from a voltage source and the other from a current source. Since the terms have negative signs, it would appear that the use of an inverting op-amp stage would be useful. Thus, one possible circuit for this application appears to be that shown in Fig. 5.28. DESIGN EXAMPLE 5.22 SOLUTION • irwin05-189-244hr2.qxd 224 22-07-2010 CHAPTER 5 9:47 Page 224 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Figure 5.28 R2 R1 Circuit used in Example 5.25. VS ± – A IS + + B Vo - The Norton equivalent circuit at the terminals A-B will provide a composite view of the op-amp’s input. Superposition can also be used in conjunction with the Norton equivalent to simplify the analysis. Using the network in Fig. 5.29a, we can determine the contribution of VS to the short-circuit current, Isc˚, which we call Isc1. Isc1 = R1 Figure 5.29 Circuits used in deriving a Norton equivalent circuit. VS ± – VS R1 R1 A Isc1 A IS R1 RTh Isc2 B B (a) (b) A B (c) In a similar manner, using Fig. 5.29b, we find that the contribution of IS to the short-circuit current is Isc2 = IS Employing superposition, the sum of these two currents yields the actual short-circuit current Isc = VS + IS R1 The Thévenin equivalent resistance at nodes A-B is obtained from the network in Fig. 5.29c as RTh = R1 The equivalent circuit is now redrawn in Fig. 5.30 where we have employed the ideal opamp conditions (i.e., Vin = 0), and the current into the op-amp terminals is zero. Since Vin is directly across RTh˚, the current in this resistor is also zero. Hence, all the current Isc will flow through R2˚, producing the voltage Vo = -R 2 c VS R2 + IS d = V - IS˚R 2 R1 R1 S A comparison of this equation with the design requirement specifies that R2 = 3 and R2 = 2000 R1 which yields R1 = 667 . Combining a 1-k and a 2-k resistor in parallel will yield the necessary 667 exactly. irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 225 SECTION 5.6 DESIGN EXAMPLES R2 Isc 0A RTh Isc Figure 5.30 The required circuit containing the Norton equivalent. 0A Vin=0 V + -+ + Vo - Fans are frequently needed to keep electronic circuits cool. They vary in size, power requirement, input voltage, and air-flow rate. In a particular application, three fans are connected in parallel to a 24-V source as shown in Fig. 5.31. A number of tests were run on this configuration, and it was found that the air flow, fan current, and input voltage are related by the following equations: FCFM = 200IF DESIGN EXAMPLE 5.23 VF = 100IF where FCFM represents the air-flow rate in cubic feet per minute, VF is the fan voltage in volts, and IF is the fan current in amperes. Note that fan current is related to fan speed, which in turn is related to air flow. A popular and inexpensive method for monitoring currents in applications where high accuracy is not critical involves placing a low-value sense resistor in series with the fan to “sense” the current by measuring the sense-resistor’s voltage. Figure 5.31 + 24 V ± – IF A trio of 24-V fans. VF - We wish to design a circuit that will measure the air flow in this three-fan system. Specifically, we want to a. determine the value of the sense resistor, placed in series with each fan, such that its voltage is 2% of the nominal 24-V fan voltage, and specify a particular 1% component that can be obtained from the Digikey Corporation (Website: www.digikey.com). b. design an op-amp circuit that will produce an output voltage proportional to total air flow, in which 1 V corresponds to 50 CFM. The fan’s voltage–current relationship specifies that each fan has a resistance of 100 . Since the voltage across the sense resistor should be 2% of 24 V, or 0.48 V, the fan current, derived from the network in Fig. 5.32, is IF = 225 24 - 0.48 = 235.2 mA 100 and the required value of the sense resistor is R sense = 0.48 = 2.04 0.2352 SOLUTION irwin05-189-244hr2.qxd 226 22-07-2010 CHAPTER 5 9:47 Page 226 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S Figure 5.32 + The equivalent circuit for one fan and its sense resistor. VF IF 100 24 V ± – + Rsense Vsense 0.48 V - The power dissipation in this component is only Psense = I 2F˚R sense = 0.11 W And thus a standard 1/4 W 2- resistor will satisfy the specifications. The op-amp circuit must be capable of adding the air-flow contributions of all three fans and scaling the result such that 1 V corresponds to 50 CFM. A summing op-amp circuit would appear to be a logical choice in this situation, and thus we select the circuit shown in Fig. 5.33 where the second stage is simply an inverter that corrects for the negative sign resulting from the summer output. In order to determine the summer’s gain, we calculate the volts/CFM at the sense resistors. For a single fan, the air flow is FCFM = 200IF = 47.04 CFM And the volts per CFM at the input to the summer are 0.48 V = 0.0102 V兾CFM 47.04 CFM Hence, the gain of the summer op-amp must be Vo 1 V兾50 CFM = = 1.96 V兾V Vsense 0.0102 V兾CFM This is a gain close to 2, and therefore we will use resistors that produce a 2:1 ratio, that is, very close to 1.96. At this point, one additional consideration must be addressed. Note that the resistors at the summer input are essentially connected in parallel with the sense resistors. To ensure that all the fan current flows in the sense resistors, we select very large values for the op-amp resistors. Let us choose R1 = R2 = R3 = 100 k and then R4 = 200 k. Figure 5.33 IF The complete air-flow measurement system. + VF R1 - R4 24 V ± – R2 R3 + Rsense Vsense - + R6 R5 + + Vo - Finally, the values for R5 and R6 can be somewhat arbitrary, as long as they are equal. If we select a value of 100 k, then only two different resistor values are needed for the entire op-amp circuit. irwin05-189-244hr4.qxd 3-09-2010 15:54 Page 227 227 PROBLEMS • SUMMARY ■ Linearity: This property requires both additivity and homogeneity. Using this property, we can determine the voltage or current somewhere in a network by assuming a specific value for the variable and then determining what source value is required to produce it. The ratio of the specified source value to that computed from the assumed value of the variable, together with the assumed value of the variable, can be used to obtain a solution. ■ In a linear network containing multiple independent sources, the principle of superposition allows us to compute any current or voltage in the network as the algebraic sum of the individual contributions of each source acting alone. ■ Superposition is a linear property and does not apply to nonlinear functions such as power. ■ Using Thévenin’s theorem, we can replace some portion of a network at a pair of terminals with a voltage source Voc in series with a resistor RTh˚. Voc is the open-circuit voltage at the terminals, and RTh is the Thévenin equivalent resistance obtained by looking into the terminals with all independent sources made zero. ■ Using Norton’s theorem, we can replace some portion of a network at a pair of terminals with a current source Isc in parallel with a resistor RTh˚. Isc is the short-circuit current at the terminals, and RTh is the Thévenin equivalent resistance. ■ Source transformation permits us to replace a voltage source V in series with a resistance R by a current source I ⫽ V/R in parallel with the resistance R. The reverse is also true. This is an interchange relationship between Thévenin and Norton equivalent circuits. ■ Maximum power transfer can be achieved by selecting the load RL to be equal to RTh found by looking into the network from the load terminals. PROBLEMS • 5.4 Find Io in the circuit in Fig. P5.4 using linearity and the 5.1 Use linearity and the assumption that Vo ⫽ 1 V to find assumption that Io = 1 mA. the actual value of Vo in Fig. P5.1. + 2 k⍀ 4 k⍀ 4 k⍀ 4 k⍀ 2 k⍀ 2 k⍀ 2 k⍀ 2 k⍀ Vo 12 k⍀ 4 k⍀ 4 mA 2 k⍀ Io Is=12 mA – Figure P5.4 Figure P5.1 5.2 Using linearity and the assumption that Io ⫽ 1 mA, find 5.5 Find Vo in the network in Fig. P5.5 using linearity and the assumption that Vo = 1 V. the actual value of Io in the network use Fig. P5.2. I1 6 k⍀ + – V 4 k⍀ 3 k⍀ s = 24 V 2 k⍀ 6 k⍀ 8V + – 3 k⍀ I3 + 3 k⍀ 3 k⍀ V2 Io – V4 3 k⍀ 6 k⍀ I3 + 3 k⍀ Vo – – Figure P5.5 5.3 Find Io in the network in Fig. P5.3 using linearity and the assumption that Io = 1 mA. 2 k⍀ 5.6 Find Io in the network in Fig. P5.6 using superposition. 6 k⍀ 2 k⍀ Io Figure P5.3 + 3 k⍀ 4 k⍀ Figure P5.2 12 mA I2 2 k⍀ 2 k⍀ 4 k⍀ 2 k⍀ 6V + – 3 k⍀ 2 k⍀ 2 mA Io Figure P5.6 3 k⍀ irwin05-189-244hr2.qxd 228 22-07-2010 CHAPTER 5 9:47 Page 228 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.7 In the network in Fig. P5.7 find Io using 5.12 Find Vo in the circuit in Fig. P5.12 using superposition. superposition. Vo + 6 k 6 k – 2 k 12 V + – 12 V 6 k 6 k 6 mA 6 mA –+ Io 2 k 2 k 2 k Figure P5.7 5.8 Find Vo in the network in Fig. P5.8 using superposition. + 3 k 12 V + – Vo Figure P5.12 – 8 k 5.13 Find Vo in the circuit in Fig. P5.13 using superposition. 6 k 2 k 2 mA + Vo 2 k 6 mA – Figure P5.8 4 mA 4 k 5.9 Find Vo in the network in Fig. P.5.9 using superposition. + 4 k Vo 2 k 2 k – 6 k 4 k 8 k Figure P5.13 6 mA 2 k 4 mA 4 k 5.14 Find Io in the circuit in Fig. P5.14 using superposition. Figure P5.9 12 V 5.10 Find Vo in the network in Fig. P5.10 using superposition. + 6 k 6 k Vo 6 k 6 k Io 6 k + – 6 mA 6 k – 6 k 12 V + – 6 k Figure P5.14 6 k 6 mA 5.15 Find Vo in the circuit in Fig. P5.15 using superposition. Figure P5.10 + 5.11 Find Io in the network in Fig. P5.11 using superposition. 6 mA 6 k 6 k 12 V + – 2 k 4 k 6 k 6 mA 4 k 12 V – + 2 k Figure P5.11 Vo 6 k – Io Figure P5.15 irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 229 229 PROBLEMS 5.16 Find Io in the circuit in Fig. P5.16 using superposition. 5.20 Use superposition to find Vo in the network in Fig. P5.20. 6 k + + – 12 V 12 V 6V 6 k – + Vo 6V – –+ +– 6 k 6 k 6 mA 6 k 6 mA 6 k 6 k 6 k Io 6 k Figure P5.20 Figure P5.16 5.21 Use superposition to find Io in the circuit in Fig. P5.21. 5.17 Use superposition to find Io in the circuit in Fig. P5.17. 12 mA + – 6 k 6 k 6 k 6 k 12 V 6 k 6 k Io Io 6 mA 12 k 4 mA + – 12 V Figure P5.21 6 k 6 k 5.22 Use superposition to find Io in the network in Fig. P5.22. 6 mA Figure P5.17 12 V + – 6 mA 5.18 Use superposition to find Io in the network in Fig. P5.18. Io 6 k 4 k 3 k 6 mA 4 mA 6 k Io 6V – + 6 k 6 k 6 k Figure P5.22 Figure P5.18 5.23 Use superposition to find Vo in the circuit in Fig. P5.23. + 5.19 Use superposition to find Vo in the circuit in Fig. P5.19. 12 V – + 6 mA 3 k 12 V – + + – 3 k + 3 k Vo 3 k 6V 3 k 3 k 3 k – – Figure P5.19 Vo 3 k Figure P5.23 irwin05-189-244hr2.qxd 230 22-07-2010 CHAPTER 5 9:47 Page 230 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.24 Find VA in Fig. P5.24 using superposition. 2 k 4 k 5.28 Find Vo in the circuit in Fig. P5.28 using superposition. 2 k 5 k 10 V 2 k + + – 10 mA VA 1 k 4 mA 6V 2 k 2 mA +– –+ – + 12 V Figure P5.24 2 k 1 k 1 k Vo – 5.25 Find I1 in Fig. P5.25 using superposition. 2 k Figure P5.28 4 k 5.29 Use superposition to find Io in the network in Fig. P5.29. 2 mA + – 8V 8 k 4 k 4 mA 2 k + – 4 k 12 V + – 6 k 2 k 16 V 4 k I1 12 k 6 mA Io Figure P5.25 Figure P5.29 5.26 Use superposition to calculate Ix in Fig. P5.26. 5.30 Use superposition to find Io in the circuit in Fig. P5.30. 12 k + – 2 k Ix 6 k 12 V 2 mA 6 k +– 4 k 3 k 12 V + – 6V – + + – 2 mA 12 V 4 k 24 V Io Figure P5.30 Figure P5.26 5.31 Use superposition to find Io in the circuit in Fig. P5.31. 4 k 5.27 Calculate Vo in Fig. P5.27 using superposition. + Vo 6 k – 6 k 6 k 12 V + – 4 k 2 k 10 k 10 k 4 mA + – 18 V 12 V 6 mA Figure P5.27 – + 2 mA 3 k Io Figure P5.31 2 k irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 231 PROBLEMS 5.32 Use Thévenin’s theorem to find Vo in the network in 231 5.36 Use Thévenin’s theorem to find Io in the network in Fig. P5.32. Fig. P5.36. 6V 12 V +– –+ 4 mA + 2 k 4 k 2 k Vo – 2 k 1 k 2 mA 1 k Figure P5.32 6 mA Io 5.33 Use Thévenin’s theorem to find Io in the circuit using Fig. P5.33. Figure P5.36 5.37 Find Io in the network in Fig. P5.37 using Thevenin’s theorem. 12 V + – 2 k 4 k 1 k 6 mA Io 4 k 2 k 1 k Io – + 4 mA 2 mA 6V 2 k Figure P5.33 Figure P5.37 5.34 Use Thévenin’s theorem to find Vo in the circuit using Fig. P5.34. Vo + 5.38 Find Vo in the circuit in Fig. P5.38 using Thévenin’s theorem. 12 V – –+ 2 k 6 mA 1 k –+ 12 V 2 k 2 k 2 k 2 k + 4 mA 1 k Vo 2 k – Figure P5.38 Figure P5.34 5.39 Find Vo in the circuit in Fig. P5.39 using Thévenin’s 5.35 Use Thévenin’s theorem to find Vo in the circuit in theorem. Fig. P5.35. + + 4 k 3 k 6V 1 k – 3 k 2 k + – + – 12 V 1 k – 6 mA Figure P5.35 Vo 2 k 6 k 6 k 2 mA Vo Figure P5.39 irwin05-189-244hr2.qxd 232 22-07-2010 CHAPTER 5 9:47 Page 232 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.40 Find Io in the circuit in Fig. P5.40 using Thévenin’s 5.43 Use Thévenin’s theorem to find Io in Fig. P5.43. theorem. 2 mA 12 V + – 4 k 2 k I0 Io 2 k 2 k 2 mA 1 k 2 k 4 k 1 k + – 12 V + – + – 8V 16 V Figure P5.40 Figure P5.43 5.41 Find Vo in the network in Fig. P5.41 using Thévenin’s 5.44 Calculate Ix in Fig. P5.44 using Thévenin’s theorem. theorem. 12 k + – 1 k 12 V 4 mA Vo 2 k 1 k 6 k + 2 k 12 V 4 k Ix + – 2 mA – – + 24 V Figure P5.41 Figure P5.44 5.42 Find Io in the network in Fig. P5.42 using Thévenin’s theorem. 5.45 Find Io in the network in Fig. P5.45 using Thévenin’s theorem. 1 k 1 k 2 k 12 V – + + – 2 k 1 k Io 1 mA 1 k 6V + – 6V 1 k Io Figure P5.42 Figure P5.45 2 mA irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 233 PROBLEMS 5.46 Find Vo in the network in Fig. P5.46 using Thévenin’s 233 5.49 Given the linear circuit in Fig. P5.49, it is known that when a 2-kΩ load is connected to the terminals A–B, the load current is 10 mA. If a 10-kΩ load is connected to the terminals, the load current is 6 mA. Find the current in a 20-kΩ load. theorem. 1 mA 1 k 0.5 k 2 mA 0.5 k +Vo – 1 k 2 mA A RTh – + ± Voc – 6V B Figure P5.46 Figure P5.49 5.50 If an 8-kΩ load is connected to the terminals of the network in Fig. P5.50, VAB 16 Ω. If a 2-kΩ load is connected to the terminals, VAB = 8 V. Find VAB if a 20-kΩ load is connected to the terminals. 5.47 Use Thévenin’s theorem to find Io in the network in Fig. P5.47. 24 V A + – 6 k Linear circuit 2 k 2 mA B Figure P5.50 2 mA 3 k 4 k 5.51 Find Io in the network in Fig. P5.51 using Norton’s Io theorem. Figure P5.47 6 k 12 V ± – 3 k 3 k 3 k 1 mA Io 5.48 Use Thévenin’s theorem to find Io in the circuit in Fig. P5.48. Figure P5.51 –+ 6 k 5.52 Use Norton’s theorem to find Io in the circuit in 4 k 18 V 6 k Fig. P5.52. 4 k 12 V 6 k 4 k 2 k +– 2 mA 1 mA 3 k 2 k Io Figure P5.48 2 k 4 mA Io Figure P5.52 4 k irwin05-189-244hr2.qxd 234 22-07-2010 CHAPTER 5 9:47 Page 234 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.53 Find Io in the network in Fig. P5.53 using Norton’s theorem. 5.57 Use Norton’s theorem to find Io in the network in Fig. P5.57. Io 2 k 6 k + – 6 k 2 k 2 mA + – 12 V 24 V 3 k 2 k 4 k 2 mA 3 k 4 k Figure P5.53 Io 5.54 Use Norton’s theorem to find Vo in the network in Fig. P5.54. Figure P5.57 2 k 5.58 Use Norton’s theorem to find Io in the circuit in Fig. P5.58. 4 k 2 mA + 24 V + – Vo –+ 4 k 2 k 4 k 18 V – 6 k 6 k 4 k 6 k Figure P5.54 5.55 Use Norton’s theorem to find Io in the circuit in Fig. P5.55. 2 mA 1 mA 3 k Io 1 k 1 k 1 k 2V 5.59 Find Vo in the network in Fig. P5.59 using Thévenin’s +– 2 mA Figure P5.58 theorem. 4 k 4V 2 + – 4 k Io +– +V – A 12 V Figure P5.55 + – VA + Vo 4 k 2 k 5.56 Find Vo in the circuit in Fig. P5.56 using Norton’s theorem. – Figure P5.59 5.60 Use Thévenin’s theorem to find Vo in the circuit in 2 k 4 mA Fig. P5.60. 1 k 6V +– 1 k 1 k Vo 1 k 2 Ix – Figure P5.56 –+ 2 k + 12 V 2 k 6V Ix –+ 1 k + Vo – Figure P5.60 irwin05-189-244hr4.qxd 3-09-2010 15:56 Page 235 235 PROBLEMS 5.65 Use Norton’s theorem to find Vo in the network in 5.61 Use Thévenin’s theorem to find Io in the circuit in Fig. P5.65. Fig. P5.61. Io 8 k⍀ 2000 Ix 6 k⍀ 2 k⍀ 4 k⍀ –+ 12 k⍀ + + 4 k⍀ 4 k⍀ Vx 12 V + 2V x – – 6V + – + – 2 k⍀ 3 mA Vo 4 k⍀ Ix Figure P5.61 – Figure P5.65 5.66 Find Vo in the circuit in Fig. P5.66 using Thévenin’s theorem. 5.62 Use Thévenin’s theorem to find Vo in the circuit in Fig. P5.62. Vo + – 12 V –+ 2 Ix 1 k⍀ 1 k⍀ 4 k⍀ + 1 k⍀ 1 k⍀ Vo Ix 2 k⍀ 4 k⍀ + 6V + – Vx Vx ––––– 2 k⍀ 1000 – + – 12 V – Figure P5.66 Figure P5.62 5.67 Find Vo in the network in Fig. P5.67 using Thévenin’s theorem. 5.63 Find Io in the circuit in Fig. P5.63 using Thévenin’s + theorem. + – 1 k⍀ 12 V Vx + – 2 k⍀ 2 mA Ix 1 k⍀ – + + 12 V – Vo 1 k⍀ 1 k⍀ 2 Vx 2 Ix 1 k⍀ Io – Figure P5.63 Figure P5.67 5.68 Use Thévenin’s theorem to find Vo in the circuit in 5.64 Find Vo in the network in Fig. P5.64 using Thévenin’s Fig. P5.68. theorem. 4 mA 1 k⍀ 3V – + 2 k⍀ 1 mA Ix Figure P5.64 – + 2 k⍀ Vo + –+ 1 k⍀ 2 Vx + 1000 Ix 6 k⍀ – 1 k⍀ 6V + Vx 1 k⍀ + Vo – – – Figure P5.68 irwin05-189-244hr3.qxd 236 2-08-2010 CHAPTER 5 16:41 Page 236 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.69 Use Thévenin’s theorem to find Vo in the circuit in 5.73 Find Vo in the circuit in Fig. P5.73 using Thévenin’s Fig. P5.69. theorem. 4 mA 12 V + – 1 k⍀ + 1 k⍀ – 2 Ix 1 k⍀ + + 2 Vx 2 mA 1 k⍀ Vx 1 k⍀ Ix Vo 1 k⍀ – – 2 mA 1 k⍀ 1 k⍀ + 1 k⍀ 12 V – + 1 k⍀ Figure P5.69 Vo – 5.70 Use Thévenin’s theorem to find Vo in the network in Figure P5.73 Fig. P5.70. 2 Vx + – 12 V 1 k⍀ + 1 k⍀ 1 k⍀ – + + 4 mA Vx 2 mA 1 k⍀ Vo 1 k⍀ 5.74 Find Vo in the network in Fig. P5.74 using Thévenin’s theorem. – – + Figure P5.70 2 k⍀ 1 k⍀ 1 k⍀ Vx – 5.71 Find Vo in the network in Fig. P5.71 using Norton’s theorem. 2 k⍀ + 12 V Vx –––– 4000 + – 2 Vx – + + 2 mA 1 k⍀ Vo 1 k⍀ – 2 k⍀ 3 k⍀ Vo Figure P5.74 + Vx 1 mA 3 k⍀ – – Figure P5.71 5.72 Find Io in the network in Fig. P5.72 using Thévenin’s theorem. 5.75 Find Vo in the network of Fig. P5.75 using Thévenin’s theorem. Io 1 k⍀ – 2 k⍀ 1 k⍀ 2 Vx + – 1 k⍀ 1 k⍀ 2 k⍀ + + – Figure P5.72 Vx – 1 k⍀ + – 2V 2 Vx ––––– 1000 Figure P5.75 + + 1 k⍀ Vo Vx 1 mA 1 k⍀ 4V 1 k⍀ 1 k⍀ 2 k⍀ – 4 mA irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 237 PROBLEMS 5.76 Use Thévenin’s theorem to find I2 in the circuit in 237 5.79 Use Thévenin’s theorem to find Vo in the network Fig. P5.76. in Fig. P5.79. 3 IA – VA 2 Ix + 2VA – 12 + + – 1 k 5A 1 k 2 mA 9 6 12 V 1 k + 1 k 7 4V + – + – 1 k 4IA 6V 1 k Ix Vo – I2 Figure P5.79 Figure P5.76 5.80 Find the Thévenin equivalent of the network in Fig. P5.80 at the terminals A-B. 1 k 2 k A 5.77 Use Thévenin’s theorem to find Vo in the circuit in Fig. P5.77. Vx 2 4 + – 1 9V B 2VA 2A 5.81 Find the Thévenin equivalent of the network in 5 Fig. P5.81 at the terminals A-B. + + 2 3IA + – 1 k Figure P5.80 8 VA Vx 1000 1 k – IA – + 1000 Ix 1 k Vo +– A – 1 k 2 k Figure P5.77 2 k Ix B Figure p5.81 5.78 Use Thévenin’s theorem to find Io in the network in Fig. P5.78. 5.82 Find the Thévenin equivalent circuit of the network in Fig. P5.82 at the terminals A–B. 3 1 k – + Vx VA 1 k 1 k –+ 6V 2 Vx 1 k 2 k + – IA 2 VA 9 6 7 4 IA 2V Io A Figure P5.78 + – 12 + – + – 5 Figure P5.82 B irwin05-189-244hr4.qxd 238 3-09-2010 CHAPTER 5 15:58 Page 238 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.83 Find the Thévenin equivalent of the network below at the terminals A-B in Fig. P5.83. 5.87 Find Vo in the network in Fig. P5.87 using source transformation. A 1 k⍀ 6 k⍀ 2 k⍀ + 3 k⍀ 2 Ix 1 k⍀ 1 k⍀ 1 k⍀ Ix B 6V + – 4 k⍀ – + Figure P5.83 Vo 12 V – 5.84 Find Io in the network in Fig. P5.84 using Thévenin’s theorem. Figure P5.87 5.88 Use source transformation to find Io in the network in 4Ix + – 1 k⍀ + Vx – Fig. P5.88. 2Vx Ix –+ 4 k⍀ 1 k⍀ 6V + – 1 k⍀ 2 k⍀ 2 mA 6 k⍀ 6V 1 k⍀ 2 mA 3V Io Io Figure P5.88 Figure P5.84 5.89 Find Vo in the network in Fig. P5.89 using source 5.85 Use source transformation to find Vo in the network transformation. in Fig. P5.85. + –+ 6 k⍀ Vo – 12 k⍀ 6V 2 mA + – 2 k⍀ + 4 k⍀ Vo – + - 3 k⍀ 3 k⍀ 4 k⍀ 24 V 3 k⍀ 2 mA 12 k⍀ 12 k⍀ – + 6V Figure P5.85 Figure P5.89 5.86 Find Io in the network in Fig. P5.86 using source transformation. 5.90 Find Io in the network in Fig. P5.90 using source transformation. 4 mA 3 k⍀ Io 6 k⍀ + – 6V Figure P5.86 4 mA 12 k⍀ 9 k⍀ 4 k⍀ 3 k⍀ 3 k⍀ 4 mA Figure P5.90 3 k⍀ 4 k⍀ 12 k⍀ 2 mA Io + – 12 V irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 239 239 PROBLEMS 5.91 Find Io in the circuit in Fig. P5.91 using source 5.94 Find Vo in the network in Fig. P5.94 using source transformation. exchange. –+ 6 k 3 k 6 k 3 k 12 V 4 k 12 V 6 k 2 mA 3 k + – + 6 k 2 k V0 4 mA 2 k – 3 k 2 k – + 6V 2 mA 2 k Io – + 12 V 2 k 1 k Figure P5.91 Figure P5.94 5.92 Use source exchange to find Io in the network in Fig. P5.92. 6V –+ 20 mA 5.95 Use source exchange to find Io in the circuit in Fig. P5.95. 6 k 8V 2 k +– 3 k 4 k 12 k 12 k 2 mA 1 k 2 k 1 k Io 2 k 12 k 6V 6 k +– Io + 24 V – + – 3 k 2 k 6 k Figure P5.95 8V Figure P5.92 5.93 Use a combination of Y–Δ transformation source transformation to find Io in the circuit in Fig. P5.93. 6 k 6V 6 k 5.96 Use source exchange to find Io in the network in Fig. P5.96. –+ Io 6 k 4 k 3 k I0 2 k 2 k 3 k 4V 4 k + – 2 mA 8 k – + 12 V 2 k 12 k 6 k 4 k 2 mA 4 k +– 6 k Figure P5.93 6V 4 k 8 k 4 k Figure P5.96 8 k 6 k 12 k irwin05-189-244hr2.qxd 240 22-07-2010 CHAPTER 5 9:47 Page 240 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.97 Use source exchange to find Io in the network in 5.100 Using source transformation, find Io in the circuit in Fig. P5.97. Fig. P5.100. 3 k 4 k 8 k 4 mA 3 k 4 k 6 k 2 k 2 mA 2 k I0 12 V 2 k + – 6 k 2 k – + 1 mA 3 k 6V 6 k 2 k 3 k Io 3 k 4 k 6 k 2 mA 2 k Figure P5.100 Figure P5.97 5.101 Use source transformation to find Io in the circuit in Fig. P5.101. 5.98 Use source transformation to find Io in the network in Fig. P5.98. 4 k 6 k 12 V 6 k 3 k 6 k – + 6 k 6 k 4 mA 12 k Io 12 V – + 2 mA 3 k 2 k Io 2 mA 3 k Figure P5.101 Figure P5.98 5.102 Using source transformation, find Io in the network in Fig. P5.102. 5.99 Using source transformation, find Vo in the circuit 2 mA in Fig. P5.99. 6 k 12 V + – 8 k 3 k 4 k 2 k + Vo 4 mA 2 k 4 k – 6 mA 12 k Io 2 mA Figure P5.99 4 k Figure P5.102 irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 241 241 PROBLEMS 5.103 Use source transformation to find Io in the circuit in 5.107 In the network in Fig. P5.107 find RL for maximum Fig. P5.103. + – 2 k power transfer and the maximum power transferred to this load. 6V 1 k 6 k 2 k 2 k +– 3 k 12 V 4 mA RL 4 k 2 mA 12 V + – Figure P5.107 4 k Io 5.108 Find RL for maximum power transfer and the Figure P5.103 maximum power that can be transferred to the load in Fig. P5.108. 5.104 Use source transformation to find Io in the circuit in Fig. P5.104. 2 mA –+ 4 k 18 V 6 k 3 k 6 k + – 6 k 2 mA 2 k 4 k 1 mA 6V RL 6 k Figure P5.108 3 k Io 5.109 Calculate the maximum power that can be transferred to RL in Fig. P5.109. Figure P5.104 5.105 Using source transformation, find Io in the circuit in 6 k Fig. P5.105. 6V 2 k 10 k 10 k RL + – + – 9V 3 mA 24 V + – 6 k 2 k Figure P5.109 2 mA 2 mA 3 k 5.110 Find the value of RL in Fig. P5.110 for maximum power 4 k transfer and the maximum power that can be dissipated in RL. Io 2 k 4 k Figure P5.105 + – 5.106 Find RL in the network in Fig. P5.106 in order to 8V RL achieve maximum power transfer. 2 k 2 k + – 12 V Figure P5.106 4 k 2 k 2 k 2 k RL + – 12 V Figure P5.110 6 k + – 6V irwin05-189-244hr2.qxd 242 22-07-2010 CHAPTER 5 15:16 Page 242 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S 5.111 Determine the value of RL in Fig. P5.111 for maximum 5.115 Find the value of RL in Fig. P5.115 for maximum power power transfer. In addition, calculate the power dissipated in RL under these conditions. transfer and the maximum power that can be transferred to RL. 8⍀ – 15 ⍀ 10 ⍀ 4A 2VA 4⍀ + RL – VA + 20 ⍀ 30 V + – + – 20 ⍀ 20 V Figure P5.111 12 V + – RL 2A Figure P5.115 5.116 Find the value of Rl in Fig. P5.116 for maximum power transfer and the maximum power that can be dissipated in RL. 5.112 Determine the value of RL in the network in Fig. 5 k⍀ 5⍀ 5⍀ P5.112 for maximum power transfer. 5 k⍀ Ix 10 ⍀ I + – 100I 12 V RL + 12 V – 2Ix + – RL Figure P5.112 Figure P5.116 5.117 Find the value of RL in Fig. P5.117 for maximum power transfer. In addition, calculate the power dissipated in RL under these conditions. 5.113 Find RL for maximum power transfer and the 4 k⍀ maximum power that can be transferred to the load in Fig. P5.113. 4 k⍀ RL 2 k⍀ 3 k⍀ + Vx 1 mA 4 Vx ––––– 1 k⍀ 1000 – 12 V RL + – 6 k⍀ 2Ix 2 mA Ix Figure P5.113 Figure P5.117 5.118 Find the value of RL in Fig. P5.118 for maximum power transfer. In addition, calculate the power dissipated in RL under these conditions. 5.114 Find the value of RL in the network in Fig. P5.114 for maximum power transfer. + Vx 2⍀ 4 Vx + – 6 k⍀ + – RL 4⍀ 2A RL 12 V + – 4⍀ Figure P5.114 3 k⍀ 3 mA Vx + 0.5V x – – Figure P5.118 2 k⍀ irwin05-189-244hr2.qxd 22-07-2010 9:47 Page 243 243 PROBLEMS 5.119 Calculate the maximum power that can be transferred 5.122 Solve the remaining problems using computational to RL in the circuit in Fig. P5.119. methods. Find Io in the network in Fig. P5.122. 6 k 4 Vx 4 12 V + – 6 k 6 mA 6 k Io 4 4 + Vx – + –+ 100 V 6 k Figure P5.122 4 RL – + – 20 V 5.123 Find Vo in the network in Fig. P5.123. –+ Figure P5.119 12 k 6V 2 mA 6 k + 4 k – + Vo – 24 V Figure P5.123 5.120 Find RL for maximum power transfer and the maximum power that can be transferred in the network in Fig. P5.120. 5.124 Find Io in the circuit in Fig. P5.124. 2 k –+ RL 4 k 1 mA Vx 2 k 2000 4 k 18 V + 2 k Vx 6 k 6 k 4 k 6 k – Figure P5.120 2 mA 1 mA 3 k Io Figure P5.124 5.121 Find the value of RL in Fig. P5.121 for maximum power transfer and the maximum power that can be dissipated in RL. 2 5.125 Find Vo in the network in Fig. P5.125. 4 IA ± – 1 8 – 2A 2 k Figure P5.121 1 k 1 k 2 k 3 IA Vx – RL VA ± – + 9V + 2 VA 2 12 V – + 2 Vx Figure P5.125 2 mA + – Io 1 k + Vo – irwin05-189-244hr2.qxd 244 22-07-2010 CHAPTER 5 9:47 Page 244 A D D I T I O N A L A N A LY S I S T E C H N I Q U E S TYPICAL PROBLEMS FOUND ON THE FE EXAM • 5PFE-1 Determine the maximum power that can be delivered 5PFE-4 What is the current I in Fig. 5PFE-4? to the load RL in the network in Fig. 5PFE-1. a. a. 2 mW b. −4 A b. 10 mW c. c. 4 mW 1 k 12 V 1 k 1 k 2 I 3 4 mA ± – 2 5PFE-5 What is the open-circuit voltage Voc at terminals a and Fig. 5PFE-2 that will achieve maximum power transfer, and determine that value of the maximum power. b of the circuit in Fig. 5PFE-5? a. 22.5 mW c. 64.3 mW 4V d. 10 V d. 121.5 mW + 8V b. 12 V b. 80.4 mW Vx a 4 2 k 12 V 20 V Figure 5PFE-4 5PFE-2 Find the value of the load RL in the network in c. 4 10 A RL 2 k Figure 5PFE-1 a. 0A d. 4 A d. 8 mW ± – 8A ± – 3 + 1 k RL ± – 12 V 2Vx – ± 2 12 A Voc b Figure 5PFE-2 Figure 5PFE-5 5PFE-3 Find the value of RL in the network in Fig. 5PFE-3 for maximum power transfer to this load. a. 12.92 Ω b. 8.22 Ω c. 6.78 Ω d. 10.53 Ω Ix 3 12 V ± – Figure 5PFE-3 12 12 2Ix RL RL irwin06-245-295hr.qxd 9-07-2010 14:27 Page 245 C H A PT E R CAPACITANCE AND INDUCTANCE 6 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Know how to use circuit models for inductors and capacitors to calculate voltage, current, and power ■ Be able to calculate stored energy for capacitors and inductors ■ Understand the concepts of continuity of current for an inductor and continuity of voltage for a capacitor ■ Be able to calculate voltages and currents for capacitors and inductors in electric circuits with dc sources ■ Know how to combine capacitors and inductors Courtesy of UPI/Brian Kersey/NewsCom A Airport Scanners To be searched or not to be searched is in series and parallel chalk etching. Millimeter wave technology emits 10,000 never the question. Air travelers demand security in the times less radio frequency than a cell phone. Backscatter skies and today’s technology makes it possible with just a technology uses high-energy x-rays as it moves through 15-to-30 second body scan instead of an intrusive pat-down clothing and other materials. In both cases, images used for that can take two to four minutes. Over 99% of airline pas- security are not retained but destroyed immediately. sengers in major airports across the nation choose to use This chapter introduces two new circuit elements: body scanners when faced with the option. Scanners can capacitors and inductors that store energy in electric and spot plastic and ceramic weapons and explosives that magnetic fields. Voltage and current relationships for evade metal detectors and could eventually replace metal these components do not follow Ohm’s law but instead detectors at the nation’s 2,000 airport checkpoints. connect voltages and currents to their derivatives and Most travelers say they welcome any measure that integrals. Capacitors and inductors are central to the enhances safety, even if it means giving up some privacy. study of alternating current circuits. They are also key Today’s new body scanners depend on millimeter wave tech- components in making body scanners work. The circuit nology or backscatter x-ray technology. The first produces an designs behind effective full-body scanners help make image that resembles a fuzzy photo negative; the second a everyone safer in the skies. What a feeling. 245 irwin06-245-295hr.qxd 246 9-07-2010 CHAPTER 6 14:27 Page 246 C A PA C I TA N C E A N D I N D U C TA N C E 6.1 Capacitors A capacitor is a circuit element that consists of two conducting surfaces separated by a nonconducting, or dielectric, material. A simplified capacitor and its electrical symbol are shown in Fig. 6.1. There are many different kinds of capacitors, and they are categorized by the type of dielectric material used between the conducting plates. Although any good insulator can serve as a dielectric, each type has characteristics that make it more suitable for particular applications. For general applications in electronic circuits (e.g., coupling between stages of amplification), the dielectric material may be paper impregnated with oil or wax, mylar, polystyrene, mica, glass, or ceramic. Ceramic dielectric capacitors constructed of barium titanates have a large capacitance-to-volume ratio because of their high dielectric constant. Mica, glass, and ceramic dielectric capacitors will operate satisfactorily at high frequencies. Aluminum electrolytic capacitors, which consist of a pair of aluminum plates separated by a moistened borax paste electrolyte, can provide high values of capacitance in small volumes. They are typically used for filtering, bypassing, and coupling, and in power supplies and motor-starting applications. Tantalum electrolytic capacitors have lower losses and more stable characteristics than those of aluminum electrolytic capacitors. Fig. 6.2 shows a variety of typical discrete capacitors. In addition to these capacitors, which we deliberately insert in a network for specific applications, stray capacitance is present any time there is a difference in potential between two conducting materials separated by a dielectric. Because this stray capacitance can cause dq i=— dt Figure 6.1 A capacitor and its electrical symbol. d A Note the use of the passive sign convention. Figure 6.2 Some typical capacitors. (Courtesy of Mark Nelms and Jo Ann Loden) + v(t) Dielectric [hint] + (a) q(t) - (b) C irwin06-245-295hr.qxd 9-07-2010 14:27 Page 247 SECTION 6.1 unwanted coupling between circuits, extreme care must be exercised in the layout of electronic systems on printed circuit boards. Capacitance is measured in coulombs per volt or farads. The unit farad (F) is named after Michael Faraday, a famous English physicist. Capacitors may be fixed or variable and typically range from thousands of microfarads (F) to a few picofarads (pF). Capacitor technology, initially driven by the modern interest in electric vehicles, is rapidly changing, however. For example, the capacitor on the left in the photograph in Fig. 6.3 is a double-layer capacitor, which is rated at 2.5 V and 100 F. An aluminum electrolytic capacitor, rated at 25 V and 68,000 F, is shown on the right in this photograph. The electrolytic capacitor can store 0.5 * 6.8 * 10-2 * 252 = 21.25 joules (J). The double-layer capacitor can store 0.5 * 100 * 2.52 = 312.5 J. Let’s connect ten of the 100-F capacitors in series for an equivalent 25-V capacitor. The energy stored in this equivalent capacitor is 3125 J. We would need to connect 147 electrolytic capacitors in parallel to store that much energy. It is interesting to calculate the dimensions of a simple equivalent capacitor consisting of two parallel plates each of area A, separated by a distance d as shown in Fig. 6.1. We learned in basic physics that the capacitance of two parallel plates of area A, separated by distance d, is C = eo A d where ´o , the permitivity of free space, is 8.85 * 10 - 12 F/m. If we assume the plates are separated by a distance in air of the thickness of one sheet of oil-impregnated paper, which is about 1.016 * 10 - 4 m, then 100 F = A8.85 * 10-12 BA 1.016 * 10-4 A = 1.148 * 109 m2 and since 1 square mile is equal to 2.59 * 106 square meters, the area is A≠443 square miles which is the area of a medium-sized city! It would now seem that the double-layer capacitor in the photograph is much more impressive than it originally appeared. This capacitor is actually constructed using a high surface area material such as powdered carbon which is adhered to a metal foil. There are literally millions of pieces of carbon employed to obtain the required surface area. Suppose now that a source is connected to the capacitor shown in Fig. 6.1; then positive charges will be transferred to one plate and negative charges to the other. The charge on the capacitor is proportional to the voltage across it such that q=Cv 6.1 where C is the proportionality factor known as the capacitance of the element in farads. The charge differential between the plates creates an electric field that stores energy. Because of the presence of the dielectric, the conduction current that flows in the wires that connect the capacitor to the remainder of the circuit cannot flow internally between the plates. However, via electromagnetic field theory it can be shown that this conduction current is equal to the displacement current that flows between the plates of the capacitor and is present any time that an electric field or voltage varies with time. Our primary interest is in the current–voltage terminal characteristics of the capacitor. Since the current is i = dq dt C A PA C I T O R S 247 Figure 6.3 A 100-F double-layer capacitor and a 68,000-F electrolytic capacitor. (Courtesy of Mark Nelms and Jo Ann Loden) irwin06-245-295hr.qxd 248 9-07-2010 CHAPTER 6 14:27 Page 248 C A PA C I TA N C E A N D I N D U C TA N C E then for a capacitor d (Cv) dt i = which for constant capacitance is i = C dv dt 6.2 Eq. (6.2) can be rewritten as dv = 1 i dt C Now integrating this expression from t=–q to some time t and assuming v(–q)=0 yields v(t) = t 1 i(x) dx C 3-q 6.3 where v(t) indicates the time dependence of the voltage. Eq. (6.3) can be expressed as two integrals, so that v(t) = t0 t 1 1 i(x) dx + i(x) dx C 3-q C 3t0 = v(t0) + t 1 i(x) dx C 3t0 6.4 where vAt0 B is the voltage due to the charge that accumulates on the capacitor from time t=–q to time t=t0 . The energy stored in the capacitor can be derived from the power that is delivered to the element. This power is given by the expression p(t) = v(t)i(t) = Cv(t) dv(t) 6.5 dt and hence the energy stored in the electric field is t wC(t) = 3-q Cv(x) dv(x) dx t dx = C v(t) = C = v(x) dv(x) = 3v(-q) 3-q v(x) dv(x) dx dx v(t) 1 2 Cv (x) 2 2 v(-q) 1 2 Cv (t) J 2 6.6 since v(t=–q)=0. The expression for the energy can also be written using Eq. (6.1) as wC(t) = 2 1 q (t) 2 C 6.7 Eqs. (6.6) and (6.7) represent the energy stored by the capacitor, which, in turn, is equal to the work done by the source to charge the capacitor. Now let’s consider the case of a dc voltage applied across a capacitor. From Eq. (6.2), we see that the current flowing through the capacitor is directly proportional to the time rate of change of the voltage across the capacitor. A dc voltage does not vary with time, so the current flowing through the capacitor is zero. We can say that a capacitor is “an open circuit to irwin06-245-295hr.qxd 9-07-2010 14:27 Page 249 SECTION 6.1 C A PA C I T O R S 249 dc” or “blocks dc.” Capacitors are often utilized to remove or filter out an unwanted dc voltage. In analyzing a circuit containing dc voltage sources and capacitors, we can replace the capacitors with an open circuit and calculate voltages and currents in the circuit using our many analysis tools. Note that the power absorbed by a capacitor, given by Eq. (6.5), is directly proportional to the time rate of change of the voltage across the capacitor. What if we had an instantaneous change in the capacitor voltage? This would correspond to dv兾dt = q and infinite power. In Chapter 1, we ruled out the possibility of any sources of infinite power. Since we only have finite power sources, the voltage across a capacitor cannot change instantaneously. This will be a particularly helpful idea in the next chapter when we encounter circuits containing switches. This idea of “continuity of voltage” for a capacitor tells us that the voltage across the capacitor just after a switch moves is the same as the voltage across the capacitor just before that switch moves. The polarity of the voltage across a capacitor being charged is shown in Fig. 6.1b. In the ideal case, the capacitor will hold the charge for an indefinite period of time, if the source is removed. If at some later time an energy-absorbing device (e.g., a flash bulb) is connected across the capacitor, a discharge current will flow from the capacitor and, therefore, the capacitor will supply its stored energy to the device. If the charge accumulated on two parallel conductors charged to 12 V is 600 pC, what is the capacitance of the parallel conductors? EXAMPLE 6.1 SOLUTION Using Eq. (6.1), we find that C = (600)A10-12 B Q = = 50 pF V 12 The voltage across a 5-F capacitor has the waveform shown in Fig. 6.4a. Determine the current waveform. EXAMPLE 6.2 SOLUTION Note that v(t) = 24 6 * 10-3 -24 = 2 * 10-3 = 0 v(t) (V) t 0 ⱕ t ⱕ 6 ms t + 96 6 ⱕ t 6 8 ms 8 ms ⱕ t Figure 6.4 i(t) (mA) Voltage and current waveforms for a 5-F capacitor. 20 24 V 0 0 6 8 t (ms) (a) 6 8 t (ms) –60 (b) irwin06-245-295hr.qxd 250 9-07-2010 CHAPTER 6 14:27 Page 250 C A PA C I TA N C E A N D I N D U C TA N C E Using Eq. (6.2), we find that i(t) = C dv (t) dt = 5 * 10-6 A4 * 103 B 0 ⱕ t ⱕ 6 ms 0 ⱕ t ⱕ 6 ms = 20 mA i(t) = 5 * 10-6 A-12 * 103 B 6 ⱕ t ⱕ 8 ms 6 ⱕ t 6 8 ms = -60 mA and 8 ms ⱕ t i(t) = 0 Therefore, the current waveform is as shown in Fig. 6.4b and i(t)=0 for t>8 ms. EXAMPLE Determine the energy stored in the electric field of the capacitor in Example 6.2 at t=6 ms. 6.3 Using Eq. (6.6), we have SOLUTION w(t) = 1 2 Cv (t) 2 At t=6 ms, w(6 ms) = 1 A5 * 10-6 B(24)2 2 = 1440 J Learning Assessment E6.1 A 10-µF capacitor has an accumulated charge of 500 nC. Determine the voltage across the capacitor. EXAMPLE 6.4 SOLUTION ANSWER: 0.05 V. The current in an initially uncharged 4-F capacitor is shown in Fig. 6.5a. Let us derive the waveforms for the voltage, power, and energy and compute the energy stored in the electric field of the capacitor at t=2 ms. The equations for the current waveform in the specific time intervals are i(t) = 16 * 10-6t 0 ⱕ t ⱕ 2 ms 2 * 10-3 = -8 * 10-6 2 ms ⱕ t ⱕ 4 ms = 0 4 ms 6 t Since v(0)=0, the equation for v(t) in the time interval 0 ⱕ t ⱕ 2 ms is v(t) = and hence, t 1 (4)A10 B 30 -6 8A10-3 Bx dx = 103t2 v(2 ms) = 103 A2 * 10-3 B = 4 mV 2 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 251 SECTION 6.1 C A PA C I T O R S 1 3 251 Voltage (mV) Current (A) 4 15 3.5 3 10 2.5 2 5 2 0 0.5 1 1.5 4 1.5 2.5 3 3.5 –5 Time (ms) 1 0.5 0 –10 0.5 1.5 (a) 2 2.5 3.5 4 Time (ms) 3.5 4 (b) Energy (pJ) 35 Power (nW) 30 60 50 40 30 25 20 15 20 10 0 –10 –20 –30 2 0.5 1 10 3.5 1.5 2.5 4 Time (ms) 3 5 0 0.5 1 1.5 2 (c) 2.5 3 Time (ms) (d) Figure 6.5 Waveforms used in Example 6.4. In the time interval 2 ms ⱕ t ⱕ 4 ms, v(t) = t 1 (4)A10 B 32A10-3B -6 - A8BA10-6 Bdx + A4B A10-3 B = -2t + 8 * 10-3 The waveform for the voltage is shown in Fig. 6.5b. Since the power is p(t)=v(t)i(t), the expression for the power in the time interval 0 ⱕ t ⱕ 2 ms is p(t)=8t3. In the time interval 2 ms ⱕ t ⱕ 4 ms, the equation for the power is p(t) = -(8)A10-6 BA-2t + 8 * 10-3 B = 16A10-6 Bt - 64A10-9 B The power waveform is shown in Fig. 6.5c. Note that during the time interval 0 ⱕ t ⱕ 2 ms, the capacitor is absorbing energy and during the interval 2 ms ⱕ t ⱕ 4 ms, it is delivering energy. The energy is given by the expression t w(t) = 3t0 p(x) dx + wAt0 B irwin06-245-295hr.qxd 252 9-07-2010 CHAPTER 6 14:27 Page 252 C A PA C I TA N C E A N D I N D U C TA N C E In the time interval 0 ⱕ t ⱕ 2 ms, t w(t) = 8x 3 dx = 2t4 30 Hence, w(2 ms)=32 pJ In the time interval 2 ⱕ t ⱕ 4 ms, t w(t) = 32 * 10-3 C A16 * 10-6 Bx - A64 * 10-9 B D dx + 32 * 10-12 = C A8 * 10-6 Bx2 - A64 * 10-9 Bx D 2 * 10-3 + 32 * 10-12 t = A8 * 10-6 Bt2 - A64 * 10-9 Bt + 128 * 10-12 From this expression we find that w(2 ms)=32 pJ and w(4 ms)=0. The energy waveform is shown in Fig. 6.5d. Learning Assessments E6.2 The voltage across a 2-F capacitor is shown in Fig. E6.2. Determine the waveform for the capacitor current. ANSWER: v(t) (V) i(t) (mA) 12 12 6 0 1 2 3 4 5 6 t (ms) –6 0 2 1 3 4 5 t (ms) Figure E6.2 E6.3 Compute the energy stored in the electric field of the capacitor in Learning Assessment ANSWER: w=144 J. E6.2 at t=2 ms. E6.4 The voltage across a 5-F capacitor is shown in Fig. E6.4. Find the waveform for the current in the capacitor. How much energy is stored in the capacitor at t=4 ms. v(t) (V) 10 5 7 9 t (ms) 1 –5 Figure E6.4 2 3 4 5 6 8 10 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 253 SECTION 6.1 C A PA C I T O R S ANSWER: 250 J. i(t) (mA) 25 4 7 t (ms) 1 2 3 5 6 9 8 10 –25 E6.5 The waveform for the current in a 1-nF capacitor is Fig. E6.5. If the capacitor has an initial voltage of –5V, determine the waveform for the capacitor voltage. How much energy is stored in the capacitor at t=6 ms? i(t) (A) 10 2 4 t (ms) 1 3 5 6 7 9 8 10 –10 Figure E6.5 ANSWER: 312.5 nJ. v(t) (V) 15 8 t (ms) –5 –15 –35 1 2 3 4 5 6 7 9 10 253 irwin06-245-295hr.qxd 254 9-07-2010 CHAPTER 6 14:27 Page 254 C A PA C I TA N C E A N D I N D U C TA N C E 6.2 An inductor is a circuit element that consists of a conducting wire usually in the form of a coil. Two typical inductors and their electrical symbol are shown in Fig. 6.6. Inductors are typically categorized by the type of core on which they are wound. For example, the core material may be air or any nonmagnetic material, iron, or ferrite. Inductors made with air or nonmagnetic materials are widely used in radio, television, and filter circuits. Iron-core inductors are used in electrical power supplies and filters. Ferrite-core inductors are widely used in high-frequency applications. Note that in contrast to the magnetic core that confines the flux, as shown in Fig. 6.6b, the flux lines for nonmagnetic inductors extend beyond the inductor itself, as illustrated in Fig. 6.6a. Like stray capacitance, stray inductance can result from any element carrying current surrounded by flux linkages. Fig. 6.7 shows a variety of typical inductors. From a historical standpoint, developments that led to the mathematical model we employ to represent the inductor are as follows. It was first shown that a current-carrying conductor would produce a magnetic field. It was later found that the magnetic field and the current that produced it were linearly related. Finally, it was shown that a changing magnetic field produced a voltage that was proportional to the time rate of change of the current that produced the magnetic field; that is, Inductors v(t) = L di(t) 6.8 dt The constant of proportionality L is called the inductance and is measured in the unit henry, named after the American inventor Joseph Henry, who discovered the relationship. As seen in Eq. (6.8), 1 henry (H) is dimensionally equal to 1 volt-second per ampere. Following the development of the mathematical equations for the capacitor, we find that the expression for the current in an inductor is i(t) = t 1 v(x) dx L 3-q 6.9 Flux lines Flux lines i(t) i(t) + + v(t) v(t) - - L i(t) (a) Figure 6.6 Two inductors and their electrical symbol Figure 6.7 Some typical inductors. (Courtesy of Mark Nelms and Jo Ann Loden) (b) (c) irwin06-245-295hr.qxd 9-07-2010 14:27 Page 255 SECTION 6.2 INDUCTORS 255 which can also be written as t 1 v(x) dx L 3t0 i(t) = iAt0 B + 6.10 The power delivered to the inductor can be used to derive the energy stored in the element. This power is equal to p(t) = v(t)i(t) di(t) = cL dt d i(t) 6.11 Therefore, the energy stored in the magnetic field is t wL(t) = 3-q cL di(x) dx d i(x) dx Following the development of Eq. (6.6), we obtain wL(t) = 1 2 Li (t) J 2 6.12 Now let’s consider the case of a dc current flowing through an inductor. From Eq. (6.8), we see that the voltage across the inductor is directly proportional to the time rate of change of the current flowing through the inductor. A dc current does not vary with time, so the voltage across the inductor is zero. We can say that an inductor is “a short circuit to dc.” In analyzing a circuit containing dc sources and inductors, we can replace any inductors with short circuits and calculate voltages and currents in the circuit using our many analysis tools. Note from Eq. (6.11) that an instantaneous change in inductor current would require infinite power. Since we don’t have any infinite power sources, the current flowing through an inductor cannot change instantaneously. This will be a particularly helpful idea in the next chapter when we encounter circuits containing switches. This idea of “continuity of current” for an inductor tells us that the current flowing through an inductor just after a switch moves is the same as the current flowing through an inductor just before that switch moves. Find the total energy stored in the circuit of Fig. 6.8a. EXAMPLE 6⍀ 9V L1=2 mH 3⍀ C1=20 F ± – L2=4 mH C2=50 F 6.5 6⍀ 3A (a) 6⍀ 9V ± – IL1 A 3⍀ + IL2 + VC1 3A - VC2 - (b) Figure 6.8 6⍀ Circuits used in Example 6.5. irwin06-245-295hr.qxd 256 9-07-2010 CHAPTER 6 14:27 Page 256 C A PA C I TA N C E A N D I N D U C TA N C E SOLUTION This circuit has only dc sources. Based on our earlier discussions about capacitors and inductors and constant sources, we can replace the capacitors with open circuits and the inductors with short circuits. The resulting circuit is shown in Fig. 6.8b. This resistive circuit can now be solved using any of the techniques we have learned in earlier chapters. If we apply KCL at node A, we get IL2 = IL1 + 3 Applying KVL around the outside of the circuit yields 6IL1 + 3IL2 + 6IL2 = 9 Solving these equations yields IL1 = -1.2 A and IL2 = 1.8 A. The voltages VC1 and VC2 can be calculated from the currents: VC1 = -6IL1 + 9 = 16.2 V VC2 = 6IL2 = 6(1.8) = 10.8 V The total energy stored in the circuit is the sum of the energy stored in the two inductors and two capacitors: 1 A2 * 10-3 B(-1.2)2 = 1.44 mJ 2 1 = A4 * 10-3 B(1.8)2 = 6.48 mJ 2 1 = A20 * 10-6 B(16.2)2 = 2.62 mJ 2 1 = A50 * 10-6 B(10.8)2 = 2.92 mJ 2 wL1 = wL2 wC1 wC2 The total stored energy is 13.46 mJ. The inductor, like the resistor and capacitor, is a passive element. The polarity of the voltage across the inductor is shown in Fig. 6.6. Practical inductors typically range from a few microhenrys to tens of henrys. From a circuit design standpoint it is important to note that inductors cannot be easily fabricated on an integrated circuit chip, and therefore chip designs typically employ only active electronic devices, resistors, and capacitors that can be easily fabricated in microcircuit form. The current in a 10-mH inductor has the waveform shown in Fig. 6.9a. Determine the voltage waveform. EXAMPLE 6.6 SOLUTION Using Eq. (6.8) and noting that i(t) = i(t) = 20 * 10-3t 2 * 10-3 -20 * 10-3t 2 * 10-3 0 ⱕ t ⱕ 2 ms 2 ⱕ t ⱕ 4 ms + 40 * 10-3 and i(t)=0 v(t) (mV) i(t) (mA) 20 100 2 Figure 6.9 Current and voltage waveforms for a 10-mH inductor. 4 ms<t 4 2 t (ms) 4 –100 (a) (b) t (ms) irwin06-245-295hr.qxd 9-07-2010 14:27 Page 257 SECTION 6.2 INDUCTORS 257 we find that v(t) = A10 * 10-3 B 20 * 10-3 0 ⱕ t ⱕ 2 ms 2 * 10-3 = 100 mV and v(t) = A10 * 10-3 B -20 * 10-3 2 ⱕ t ⱕ 4 ms 2 * 10-3 = -100 mV and v(t)=0 for t>4 ms. Therefore, the voltage waveform is shown in Fig. 6.9b. EXAMPLE The current in a 2-mH inductor is 6.7 i(t)=2 sin 377t A Determine the voltage across the inductor and the energy stored in the inductor. SOLUTION From Eq. (6.8), we have v(t) = L di(t) dt = A2 * 10-3 B d (2 sin 377t) dt = 1.508 cos 377t V and from Eq. (6.12), 1 2 Li (t) 2 1 = A2 * 10-3 B(2 sin 377t)2 2 = 0.004 sin2 377t J wL(t) = The voltage across a 200-mH inductor is given by the expression v(t) = (1 - 3t)e mV = 0 6.8 t 6 0 Let us derive the waveforms for the current, energy, and power. The waveform for the voltage is shown in Fig. 6.10a. The current is derived from Eq. (6.10) as i(t) = EXAMPLE tⱖ0 -3t t 103 (1 - 3x)e-3x dx 200 30 t t = 5e 30 = 5e t e-3x t e-3x 2 - 3c(3x + 1) d f -3 0 9 0 e-3x dx - 3 = 5te-3t mA t ⱖ 0 = 0 t 6 0 30 xe-3x dx f SOLUTION irwin06-245-295hr.qxd 258 9-07-2010 CHAPTER 6 14:27 Page 258 C A PA C I TA N C E A N D I N D U C TA N C E A plot of the current waveform is shown in Fig. 6.10b. The power is given by the expression p(t) = v(t)i(t) = 5t(1 - 3t)e-6t W tⱖ0 = 0 t 6 0 The equation for the power is plotted in Fig. 6.10c. The expression for the energy is w(t) = 1 2 Li (t) 2 = 2.5t2e-6t J tⱖ0 = 0 t 6 0 This equation is plotted in Fig. 6.10d. Voltage (mV) Current (mA) 0.7 1.0 0.6 0.8 0.5 0.6 0.4 0.4 0.3 0.2 0 0.2 0.5 1 0.1 1.5 2 2.5 3 3.5 –0.2 Time (s) 0 0.5 1 1.5 (a) 2 2.5 3 3.5 Time (s) 2.5 Time (s) (b) Energy (nJ) Power (W) 40 0.2 35 0.15 30 25 0.1 20 0.05 15 1 0 0.5 1.5 –0.05 2 2.5 10 Time (s) 5 –0.1 0 (c) Figure 6.10 Waveforms used in Example 6.8. 0.5 1 1.5 (d) 2 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 259 SECTION 6.2 INDUCTORS 259 Learning Assessments E6.6 The current in a 5-mH inductor has the waveform shown in Fig. E6.6. Compute the waveform for the inductor voltage. ANSWER: i(t) (mA) v(t) (mV) 20 100 10 0 1 2 3 4 t (ms) –50 0 1 2 3 4 t (ms) Figure E6.6 E6.7 Compute the energy stored in the magnetic field of the inductor in Learning Assesment ANSWER: W=562.5 nJ. E6.6 at t=1.5 ms. E6.8 The current in a 2-H inductor is shown in Fig. E6.8. Find the waveform for the inductor voltage. How much energy is stored in the inductor at t=3 ms? i(t) (mA) 5 2 6 t (ms) 1 3 4 5 7 8 9 10 11 12 –5 –10 Figure E6.8 ANSWER: 25 J. v(t) (V) 10 2 4 9 6 12 t (ms) 1 –3.33 –5 3 5 7 8 10 11 irwin06-245-295hr.qxd 260 22-07-2010 CHAPTER 6 15:18 Page 260 C A PA C I TA N C E A N D I N D U C TA N C E E6.9 The voltage across a 0.1-H inductor is shown in Fig. E6.9. Compute the waveform for the current in the inductor if i(0)=0.1A. How much energy is stored in the inductor at t=7 ms? v(t) (V) 10 5 2 4 6 9 t (ms) 1 3 5 7 8 10 –5 –10 Figure E6.9 ANSWER: 1.125 mJ. i(t) (A) 0.2 0.1 7 t (ms) 1 2 3 4 5 6 8 9 10 –0.1 –0.25 ANSWER: 0.72 J, 0.5 J. E6.10 Find the energy stored in the capacitor and inductor in Fig. E6.10. 2 mA 10 nF 1H 6 k⍀ 8 mA 3 k⍀ 2 k⍀ 12 V Figure E6.10 + - irwin06-245-295hr.qxd 9-07-2010 14:27 Page 261 SECTION 6.2 CAPACITOR AND INDUCTOR SPECIFICATIONS There are a couple of important parameters that are used to specify capacitors and inductors. In the case of capacitors, the capacitance value, working voltage, and tolerance are issues that must be considered in their application. Standard capacitor values range from a few pF to about 50 mF. Capacitors larger than 1 F are available but will not be discussed here. Table 6.1 is a list of standard capacitor values, which are typically given in picofarads or microfarads. Although both smaller and larger ratings are available, the standard working voltage, or dc voltage rating, is typically between 6.3 V and 500 V. Manufacturers specify this working voltage since it is critical to keep the applied voltage below the breakdown point of the dielectric. Tolerance is an adjunct to the capacitance value and is usually listed as a percentage of the nominal value. Standard tolerance values are ; 5%, ; 10%, and ; 20%. Occasionally, tolerances for single-digit pF capacitors are listed in pF. For example, 5 pF ; 0.25 pF. TABLE 6.1 Standard capacitor values pF pF pF pF F F F F F F F 1 10 12 15 18 20 22 27 33 39 47 51 56 68 82 100 120 150 180 200 220 270 330 390 470 510 560 680 820 1000 1200 1500 1800 2000 2200 2700 3300 3900 4700 5100 5600 6800 8200 0.010 0.012 0.015 0.018 0.020 0.022 0.027 0.033 0.039 0.047 0.051 0.056 0.068 0.082 0.10 0.12 0.15 0.18 0.20 0.22 0.27 0.33 0.39 0.47 0.51 0.56 0.68 0.82 1.0 1.2 1.5 1.8 2.0 2.2 2.7 3.3 3.9 4.7 5.1 5.6 6.8 8.2 10 12 15 18 20 22 27 33 39 47 51 56 68 82 100 120 150 180 200 220 270 330 390 470 510 560 680 820 1000 1200 1500 1800 2000 2200 2700 3300 3900 4700 5100 5600 6800 8200 10,000 12,000 15,000 18,000 20,000 22,000 27,000 33,000 39,000 47,000 51,000 56,000 68,000 82,000 1.5 2 3 4 5 6 7 8 9 The two principal inductor specifications are inductance and resistance. Standard commercial inductances range from about 1 nH to around 100 mH. Larger inductances can, of course, be custom built for a price. Table 6.2 lists the standard inductor values. The current rating for inductors typically extends from a few dozen mA’s to about 1 A. Tolerances are typically 5% or 10% of the specified value. TABLE 6.2 Standard inductor values nH nH nH H H H mH mH mH 1 1.2 1.5 1.8 2 2.2 2.7 3 4 5 6 7 8 9 10 12 15 18 20 22 27 33 39 47 51 56 68 82 100 120 150 180 200 220 270 330 390 470 510 560 680 820 1.0 1.2 1.5 1.8 2.0 2.2 2.7 3.3 3.9 4.7 5.1 5.6 6.8 8.2 10 12 15 18 20 22 27 33 39 47 51 56 68 82 100 120 150 180 200 220 270 330 390 470 510 560 680 820 1.0 1.2 1.5 1.8 2.0 2.2 2.7 3.3 3.9 4.7 5.1 5.6 6.8 8.2 10 12 15 18 20 22 27 33 39 47 51 56 68 82 100 INDUCTORS 261 irwin06-245-295hr.qxd 262 9-07-2010 CHAPTER 6 14:27 Page 262 C A PA C I TA N C E A N D I N D U C TA N C E As indicated in Chapter 2, wire-wound resistors are simply coils of wire, and therefore it is only logical that inductors will have some resistance. The major difference between wirewound resistors and inductors is the wire material. High-resistance materials such as Nichrome are used in resistors, and low-resistance copper is used in inductors. The resistance of the copper wire is dependent on the length and diameter of the wire. Table 6.3 lists the American Wire Gauge (AWG) standard wire diameters and the resulting resistance per foot for copper wire. TABLE 6.3 Resistance per foot of solid copper wire AWG No. 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 EXAMPLE 6.9 SOLUTION EXAMPLE 6.10 SOLUTION Diameter (in.) 0.0808 0.0641 0.0508 0.0400 0.0320 0.0253 0.0201 0.0159 0.0126 0.0100 0.0080 0.0063 0.0049 0.0039 0.0031 m⍀/ft 1.59 2.54 4.06 6.50 10.4 16.5 26.2 41.6 66.2 105 167 267 428 684 1094 We wish to find the possible range of capacitance values for a 51-mF capacitor that has a tolerance of 20%. The minimum capacitor value is 0.8C=40.8 mF, and the maximum capacitor value is 1.2C=61.2 mF. The capacitor in Fig. 6.11a is a 100-nF capacitor with a tolerance of 20%. If the voltage waveform is as shown in Fig. 6.11b, let us graph the current waveform for the minimum and maximum capacitor values. The maximum capacitor value is 1.2C=120 nF, and the minimum capacitor value is 0.8C=80 nF. The maximum and minimum capacitor currents, obtained from the equation i(t) = C are shown in Fig. 6.11c. dv(t) dt irwin06-245-295hr.qxd 9-07-2010 14:27 Page 263 SECTION 6.2 INDUCTORS 263 4 3 v(t) (V) 2 i(t) 1 0 –1 –2 –3 v(t) ± – C –4 0 1 2 6 7 (b) v(t) i(t) at Cmin i(t) at Cmax 4 3 Voltage (V) 5 800 600 2 400 1 200 0 0 –1 –200 –2 –400 –3 –600 Current (mA) (a) 3 4 Time (s) –800 –4 0 1 2 3 4 5 6 7 Time (s) Figure 6.11 (c) Circuit and graphs used in Example 6.10. The inductor in Fig. 6.12a is a 100-H inductor with a tolerance of 10%. If the current waveform is as shown in Fig. 6.12b, let us graph the voltage waveform for the minimum and maximum inductor values. The maximum inductor value is 1.1L=110 H, and the minimum inductor value is 0.9L = 90 H. The maximum and minimum inductor voltages, obtained from the equation v(t) = L are shown in Fig. 6.12c. di(t) dt EXAMPLE 6.11 SOLUTION irwin06-245-295hr.qxd 264 9-07-2010 CHAPTER 6 14:27 Page 264 C A PA C I TA N C E A N D I N D U C TA N C E Figure 6.12 Circuit and graphs used in Example 6.11. + i(t) L v(t) - (a) 150 100 100 1 50 50 0 0 –1 –50 –2 –100 –3 0 –50 –100 –150 –150 0 10 20 30 40 50 0 60 10 20 40 50 –4 60 Time (s) Time (s) (b) 6.3 30 2 v(t) (V) v(t) at Lmax 150 i(t) (mA) i(t) (mA) i(t) v(t) at Lmin (c) S E R I E S C A PA C I TO R S If a number of capacitors are connected in series, their equivalent capacitance can be calculated using KVL. Consider the circuit shown in Fig. 6.13a. For this circuit Capacitor and Inductor Combinations v(t)=v1(t)+v2(t)+v3(t)+p+vN(t) vi(t) = t 1 i(t) dt + vi At0 B Ci 3t0 6.14 v (t) v (t) v (t) i(t) +1 - +2 - +3 - Figure 6.13 Equivalent circuit for N series-connected capacitors. 6.13 but + v(t) - C1 vN(t) - + CN (a) C2 C3 i(t) + v(t) - (b) CS irwin06-245-295hr.qxd 9-07-2010 14:27 Page 265 SECTION 6.3 C A PA C I T O R A N D I N D U C T O R C O M B I N AT I O N S 265 Therefore, Eq. (6.13) can be written as follows using Eq. (6.14): N t N 1 b i(t) dt + a vi At 0 B v(t) = a a i = 1 Ci 3 t0 i=1 = t 1 i(t) dt + vAt0 B CS 3t0 6.15 6.16 where N vAt0 B = a vi At0 B i=1 and N 1 1 1 1 1 = a = + + p + CS C C C C i 1 2 N i=1 [hint] 6.17 Capacitors in series combine like resistors in parallel. Thus, the circuit in Fig. 6.13b is equivalent to that in Fig. 6.13a under the conditions stated previously. It is also important to note that since the same current flows in each of the series capacitors, each capacitor gains the same charge in the same time period. The voltage across each capacitor will depend on this charge and the capacitance of the element. Determine the equivalent capacitance and the initial voltage for the circuit shown in Fig. 6.14. Note that these capacitors must have been charged before they were connected in series or else the charge of each would be equal and the voltages would be in the same direction. The equivalent capacitance is 1 1 1 1 = + + CS 2 3 6 where all capacitance values are in microfarads. Therefore, CS=1 F and, as seen from the figure, vAt0 B = -3 V. Note that the total energy stored in the circuit is wAt0 B = 1 C2 * 10-6(2)2 + 3 * 10-6(-4)2 + 6 * 10-6(-1)2 D 2 = 31 J However, the energy recoverable at the terminals is wC At0 B = = 1 C v2(t) 2 S 1 C 1 * 10-6(-3)2 D 2 = 4.5 J 2V + + v(t) - 2 F 3 F 6 F + 1V Figure 6.14 4V + Circuit containing multiple capacitors with initial voltages. EXAMPLE SOLUTION 6.12 irwin06-245-295hr.qxd 266 9-07-2010 CHAPTER 6 14:27 Page 266 C A PA C I TA N C E A N D I N D U C TA N C E EXAMPLE 6.13 SOLUTION Two previously uncharged capacitors are connected in series and then charged with a 12-V source. One capacitor is 30 F and the other is unknown. If the voltage across the 30-F capacitor is 8 V, find the capacitance of the unknown capacitor. The charge on the 30-F capacitor is Q=CV=(30 F)(8 V)=240 C Since the same current flows in each of the series capacitors, each capacitor gains the same charge in the same time period: 240 C Q = = 60 F V 4V C = To determine the equivalent capacitance of N capacitors connected in parallel, we employ KCL. As can be seen from Fig. 6.15a, PA R A L L E L C A PA C I TO R S i(t) = i1(t) + i2(t) + i3(t) + p + iN(t) dv(t) = C1 dt dv(t) N dv(t) i=1 dt = a a Ci b = Cp [hint] + C2 dt + C3 dv(t) dt 6.18 + p + CN dv(t) dt dv(t) 6.19 dt where Cp=C1+C2+C3+p+CN Capacitors in parallel combine like resistors in series. i(t) i(t) Figure 6.15 Equivalent circuit for N capacitors connected in parallel. + i1(t) i2(t) i3(t) iN(t) C1 C2 C3 CN v(t) + v(t) - (b) Determine the equivalent capacitance at terminals A-B of the circuit shown in Fig. 6.16. 6.14 SOLUTION Figure 6.16 Circuit containing multiple capacitors in parallel. Cp (a) EXAMPLE 6.20 Cp = 15 F A + v(t) - B 4 F 6 F 2 F 3 F irwin06-245-295hr.qxd 9-07-2010 14:27 Page 267 SECTION 6.3 C A PA C I T O R A N D I N D U C T O R C O M B I N AT I O N S 267 Learning Assessments E6.11 Two initially uncharged capacitors are connected as shown in Fig. E6.11. After a period ANSWER: C1=4 F. of time, the voltage reaches the value shown. Determine the value of C1. + C1 24 V + 12 F 6V - - Figure E6.11 E6.12 Compute the equivalent capacitance of the network in Fig. E6.12. ANSWER: Ceq=1.5 F. 3 F 2 F 4 F Ceq 2 F 3 F 12 F Figure E6.12 E6.13 Determine CT in Fig. E6.13. ANSWER: 1.667 F. 6 F A 4 F 8 F 5 F CT 3 F B 6 F 2 F 10 F 6 F 6 F Figure E6.13 If N inductors are connected in series, the equivalent inductance of the combination can be determined as follows. Referring to Fig. 6.17a and using KVL, we see that S E R I E S I N D U C TO R S v(t)=v1(t)+v2(t)+v3(t)+p+vN(t) 6.21 and therefore, di(t) v(t) = L1 dt + L2 di(t) N di(t) i=1 dt = a a Li b = LS dt + L3 di(t) dt + p + LN di(t) dt 6.22 di(t) dt 6.23 irwin06-245-295hr.qxd 268 9-07-2010 CHAPTER 6 14:27 Page 268 C A PA C I TA N C E A N D I N D U C TA N C E where [hint] N LS = a Li = L1 + L2 + p + LN Inductors in series combine like resistors in series. 6.24 i=1 Therefore, under this condition the network in Fig. 6.17b is equivalent to that in Fig. 6.17a. i(t) Figure 6.17 Equivalent circuit for N series-connected inductors. + +v1(t)- +v2(t)- +v3(t)- L2 L3 L1 i(t) + v(t) v(t) - - LN - LS + vN(t) (a) (b) EXAMPLE Find the equivalent inductance of the circuit shown in Fig. 6.18. 6.15 SOLUTION The equivalent inductance of the circuit shown in Fig. 6.18 is LS=1H+2H+4H =7H Figure 6.18 1H Circuit containing multiple inductors. 2H + v(t) 4H - PA R A L L E L I N D U C TO R S Consider the circuit shown in Fig. 6.19a, which contains N parallel inductors. Using KCL, we can write i(t)=i1(t)+i2(t)+i3(t)+p+iN(t) 6.25 However, ij(t) = t 1 v(x) dx + ij At0 B Lj 3t0 6.26 Substituting this expression into Eq. (6.25) yields N t N 1 b v(x) dx + a i j At 0 B i(t) = a a j = 1 Lj 3 t0 j=1 = t 1 v(x) dx + iAt0 B Lp 3t0 6.27 6.28 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 269 SECTION 6.3 C A PA C I T O R A N D I N D U C T O R C O M B I N AT I O N S where 269 [hint] 1 1 1 1 1 = + + + p + Lp L1 L2 L3 LN 6.29 Inductors in parallel combine like resistors in parallel. and iAt0 B is equal to the current in Lp at t=t0 . Thus, the circuit in Fig. 6.19b is equivalent to that in Fig. 6.19a under the conditions stated previously. i(t) i(t) + i1(t) i2(t) i3(t) iN(t) + v(t) L1 L2 L3 LN v(t) - Figure 6.19 Lp Equivalent circuits for N inductors connected in parallel. (a) (b) Determine the equivalent inductance and the initial current for the circuit shown in Fig. 6.20. EXAMPLE 6.16 SOLUTION The equivalent inductance is 1 1 1 1 = + + Lp 12 6 4 where all inductance values are in millihenrys: Lp=2 mH and the initial current is iAt0 B = -1 A. i(t) Figure 6.20 + 3A 6A 2A v(t) 12 mH 6 mH 4 mH Circuit containing multiple inductors with initial currents. - The previous material indicates that capacitors combine like conductances, whereas inductances combine like resistances. Learning Assessment E6.14 Determine the equivalent inductance of the network in Fig. E6.14 if all inductors are 6 mH. Leq Figure E6.14 ANSWER: 9.429 mH. irwin06-245-295hr.qxd 270 9-07-2010 CHAPTER 6 14:27 Page 270 C A PA C I TA N C E A N D I N D U C TA N C E E6.15 Find LT in Fig. E6.15. ANSWER: 5 mH. 2 mH 2 mH Figure E6.15 6 mH 2 mH 4 mH LT 5 mH A B 3 mH 12 mH 4 mH 2 mH C H I P C A PA C I TO R S In Chapter 2, we briefly discussed the resistors that are used in modern electronic manufacturing. An example of these surface mount devices was shown in Fig. 2.41, together with some typical chip capacitors. As we will indicate in the material that follows, modern electronics employs primarily resistors and capacitors and avoids the use of inductors when possible. Surface-mounted chip capacitors account for the majority of capacitors used in electronics assembly today. These capacitors have a large range of sizes, from as small as 10 mils on a side up to 250 mils on a side. All ceramic chip capacitors consist of a ceramic dielectric layer between metal plates. The properties of the ceramic and metal layers determine the type of capacitor, its capacitance, and reliability. A cut-away view of a standard chip capacitor is shown in Fig. 6.21. The inner metal electrodes are alternately connected to the opposing sides of the chip where metal terminators are added. These terminators not only make connection to the inner electrodes, but also provide a solder base for attaching these chips to printed Figure 6.21 Ceramic dielectric Cross section of a multilayer ceramic chip capacitor. Tin Nickel Copper Inner electrodes (Ni/Cu) circuit boards. The number of alternating layers, the spacing between them, along with the dielectric constant of the ceramic material, will determine the capacitance value. We indicated earlier that resistors are normally manufactured in standard sizes with specific power ratings. Chip capacitors are also manufactured in this manner, and Table 6.4 provides a partial listing of these devices. The standard sizes of chip capacitors are shown in Table 6.4. irwin06-245-295hr.qxd 9-07-2010 14:27 Page 271 SECTION 6.3 C A PA C I T O R A N D I N D U C T O R C O M B I N AT I O N S 271 TABLE 6.4 Ceramic chip capacitor standard sizes Size Code Size (Mils) Power Rating (Watts) 0201 0402 0603 0805 1206 2010 2512 20 ⫻ 10 40 ⫻ 20 60 ⫻ 30 80 ⫻ 50 120 ⫻ 60 200 ⫻ 100 250 ⫻ 120 1/20 1/16 1/10 1/8 1/4 1/2 1 C H I P I N D U C TO R S A chip inductor consists of a miniature ceramic substrate with either a wire wrapped around it or a thin film deposited and patterned to form a coil. They can be encapsulated or molded with a material to protect the wire from the elements or left unprotected. Chip inductors are supplied in a variety of types and values, with three typical configurations that conform to the standard “chip” package widely utilized in the printed circuit board (PCB) industry. The first type is the precision chip inductor where copper is deposited onto the ceramic and patterned to form a coil, as shown in Fig. 6.22. L W T Copper (Cu) Termination Base 2 µm Nickel (Ni) Barrier Etched Copper Coil E Terminal Electrode 3 µm Tin (Sn) Outerplating Ferrite Internal Medium Alumina Substrate Figure 6.22 Figure 6.23 Precision chip inductor cross section. Ferrite chip inductor cross section The second type is a ferrite chip inductor, which uses a series of coil patterns stacked between ferrite layers to form a multiplayer coil as shown in Fig. 6.23. The third type is a wire-wound open frame in which a wire is wound around a ceramic substrate to form the inductor coil. The completed structure is shown in Fig. 6.24. Each of these configurations displays different characteristics, with the wire-wound type providing the highest inductance values (10 nH⫺4.7 uH)—and reasonable tolerances (1–2%). The ferrite chip inductor gives a wide range of values (47 nH⫺33 uH) but has tolerances in the 5% range. The precision chip inductor has low inductance values (1⫺100 nH) but very good tolerances (⫹/⫺0.1 nH). C Figure 6.24 E G F A F B K D Wire-wound chip inductor cross section irwin06-245-295hr.qxd 272 9-07-2010 CHAPTER 6 14:27 Page 272 C A PA C I TA N C E A N D I N D U C TA N C E 6.4 RC Operational Amplifier Circuits [hint] Two very important RC op-amp circuits are the differentiator and the integrator. These circuits are derived from the circuit for an inverting op-amp by replacing the resistors R1 and R2 , respectively, by a capacitor. Consider, for example, the circuit shown in Fig. 6.25a. The circuit equations are C1 vo - vd = iAv1 - v- B + dt R2 However, v–=0 and i–=0. Therefore, The properties of the ideal op-amp are v⫹ ⫽ v⫺ and i⫹ ⫽ i⫺ ⫽ 0. vo(t) = -R2 C1 dv1(t) 6.30 dt C2 R2 C1 v– i– v± i± v1(t) ± – + R1 v – i– v± i± + vo v1(t) ± – + + vo - - (a) (b) Figure 6.25 Differentiator and integrator operational amplifier circuits. Thus, the output of the op-amp circuit is proportional to the derivative of the input. The circuit equations for the op-amp configuration in Fig. 6.25b are v1 - vd + C2 Avo - v- B = iR1 dt but since v–=0 and i–=0, the equation reduces to v1 dvo = -C2 R1 dt or vo(t) = = t -1 v (x) dx R1 C2 3-q 1 t -1 v (x) dx + vo(0) R1 C2 30 1 6.31 If the capacitor is initially discharged, then vo(0)=0; hence, vo(t) = t -1 v (x) dx R1 C2 30 1 6.32 Thus, the output voltage of the op-amp circuit is proportional to the integral of the input voltage. irwin06-245-295hr.qxd 9-07-2010 14:27 Page 273 SECTION 6.4 R C O P E R AT I O N A L A M P L I F I E R C I R C U I T S EXAMPLE The waveform in Fig. 6.26a is applied at the input of the differentiator circuit shown in Fig. 6.25a. If R2=1 k⍀ and C1=2 F, determine the waveform at the output of the op-amp. 6.17 SOLUTION Using Eq. (6.30), we find that the op-amp output is vo(t) = -R2 C1 273 dv1(t) dt dv1(t) = -(2)10-3 dt dv1(t)/dt=(2)103 for 0 ⱕ t<5 ms, and therefore, 0 ⱕ t<5 ms vo(t)=–4 V dv1(t)/dt=–(2)103 for 5 ⱕ t<10 ms, and therefore, 5 ⱕ t<10 ms vo(t)=4 V Hence, the output waveform of the differentiator is shown in Fig. 6.26b. v1(t) (V) vo(t) (V) Figure 6.26 Input and output waveforms for a differentiator circuit. 10 +4 0 5 10 t (ms) (a) 0 5 10 –4 (b) If the integrator shown in Fig. 6.25b has the parameters R1=5 k⍀ and C2=0.2 F, determine the waveform at the op-amp output if the input waveform is given as in Fig. 6.27a and the capacitor is initially discharged. t -1 v (x) dx R1 C2 30 1 which with the given circuit parameters is t vo(t) = -103 30 v1(x) dx In the interval 0 ⱕ t<0.1 s, v1(t)=20 mV. Hence, vo(t) = -103(20)10-3t EXAMPLE 6.18 SOLUTION The integrator output is given by the expression vo(t) = t (ms) 0 ⱕ t 6 0.1 s = -20t At t=0.1 s, vo(t)=–2 V. In the interval from 0.1 to 0.2 s, the integrator produces a positive slope output of 20t from vo(0.1)=–2 V to vo(0.2)=0 V. This waveform from t=0 to t=0.2 s is repeated in the interval t=0.2 to t=0.4 s, and therefore, the output waveform is shown in Fig. 6.27b. irwin06-245-295hr.qxd 274 9-07-2010 CHAPTER 6 14:27 Page 274 C A PA C I TA N C E A N D I N D U C TA N C E v1(t) (mV) vo(t) (V) 20 0 0 0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4 t (s) t (s) –20 –2 (a) Figure 6.27 (b) Input and output waveforms for an integrator circuit. Learning Assessment E6.16 The waveform in Fig. E6.16 is applied to the input terminals of the op-amp differentiator circuit. Determine the differentiator output waveform if the op-amp circuit parameters are C1=2 F and R2=2 ⍀. ANSWER: v1(t) (V) vo(t) (V) 6 24 1 2 3 0 1 2 3 4 t (s) 4 t (s) –24 Figure E6.16 6.5 Application Examples APPLICATION EXAMPLE 6.19 In integrated circuits, wires carrying high-speed signals are closely spaced as shown by the micrograph in Fig. 6.28. As a result, a signal on one conductor can “mysteriously” appear on a different conductor. This phenomenon is called crosstalk. Let us examine this condition and propose some methods for reducing it. irwin06-245-295hr.qxd 9-07-2010 14:27 Page 275 SECTION 6.5 A P P L I C AT I O N E X A M P L E S 275 Figure 6.28 SEM Image (Tom Way/ Ginger Conly. Courtesy of International Business Machines Corporation. Unauthorized use not permitted.) The origin of crosstalk is capacitance. In particular, it is undesired capacitance, often called parasitic capacitance, that exists between wires that are closely spaced. The simple model in Fig. 6.29 can be used to investigate crosstalk between two long parallel wires. A signal is applied to wire 1. Capacitances C1 and C2 are the parasitic capacitances of the conductors with respect to ground, while C12 is the capacitance between the conductors. Recall that we introduced the capacitor as two closely spaced conducting plates. If we stretch those plates into thin wires, certainly the geometry of the conductors would change and thus the amount of capacitance. However, we should still expect some capacitance between the wires. SOLUTION Wire 1 Figure 6.29 Wire 2 A simple model for investigating crosstalk. v2(t) i12(t) v1(t) ± – C1 C12 i2(t) C2 In order to quantify the level of crosstalk, we want to know how much of the voltage on wire 1 appears on wire 2. A nodal analysis at wire 2 yields i12(t) = C12 c dv1(t) dt dv2(t) - dt d = i2(t) = C2 c Solving for dv2(t)兾dt, we find that dv2(t) dt = c dv1(t) C12 d C12 + C2 dt Integrating both sides of this equation yields v2(t) = c C12 d v (t) C12 + C2 1 dv2(t) dt d irwin06-245-295hr.qxd 276 9-07-2010 CHAPTER 6 14:27 Page 276 C A PA C I TA N C E A N D I N D U C TA N C E Note that it is a simple capacitance ratio that determines how effectively v1(t) is “coupled” into wire 2. Clearly, ensuring that C12 is much less than C2 is the key to controlling crosstalk. How is this done? First, we can make C12 as small as possible by increasing the spacing between wires. Second, we can increase C2 by putting it closer to the ground wiring. Unfortunately, the first option takes up more real estate, and the second one slows down the voltage signals in wire 1. At this point, we seem to have a typical engineering tradeoff: to improve one criterion, that is, decreased crosstalk, we must sacrifice another, space or speed. One way to address the space issue would be to insert a ground connection between the signal-carrying wires as shown in Fig. 6.30. However, any advantage achieved with grounded wires must be traded off against the increase in space, since inserting grounded wires between adjacent conductors would nearly double the width consumed without them. Wire 1 Figure 6.30 Ground wire Wire 2 Use of a ground wire in the crosstalk model. v1(t) ± – C1G + C2G C1 v2(t) C2 - Redrawing the circuit in Fig. 6.31 immediately indicates that wires 1 and 2 are now electrically isolated and there should be no crosstalk whatsoever—a situation that is highly unlikely. Thus, we are prompted to ask the question, “Is our model accurate enough to model crosstalk?” A more accurate model for the crosstalk reduction scheme is shown in Fig. 6.32 where the capacitance between signal wires 1 and 2 is no longer ignored. Once again, we will determine the amount of crosstalk by examining the ratio v2(t)兾v1(t). Employing nodal analysis at wire 2 in the circuit in Fig. 6.33 yields i12(t) = C12 c dv1(t) dv2(t) - dt dt d = i2(t) = AC2 + C2G B c Wire 1 Figure 6.31 dv2(t) dt d Wire 2 Electrical isolation using a ground wire in crosstalk model. + v1(t) ± – C1G C1 C2G C2 Ground wire v2(t) - Solving for dv2(t)兾dt, we obtain dv2(t) dt = c C12 dv1(t) C12 d + C2 + C2G dt Integrating both sides of this equation yields v2(t) = c C12 d v (t) C12 + C2 + C2G 1 Note that this result is very similar to our earlier result with the addition of the C2G term. Two benefits in this situation reduce crosstalk. First, C12 is smaller because adding the ground wire moves wires 1 and 2 farther apart. Second, C2G makes the denominator of the crosstalk equation bigger. If we assume that C2G = C2 and that C12 has been halved by the extra spacing, we can expect the crosstalk to be reduced by a factor of roughly 4. irwin06-245-295hr.qxd 9-07-2010 14:27 Page 277 SECTION 6.5 A P P L I C AT I O N E X A M P L E S 277 Figure 6.32 C12 A more accurate crosstalk model. Wire 1 Ground wire C1G v1(t) ± – Wire 2 + C2G C1 v2(t) C2 - Figure 6.33 C12 A redrawn version of the more accurate crosstalk model. i12(t) Wire 1 Wire 2 i2(t) v1(t) ± – C2 v2(t) C1G C2G C1 + - An excellent example of capacitor operation is the memory inside a personal computer. This memory, called dynamic random access memory (DRAM), contains as many as 4 billion data storage sites called cells (circa 2007). Expect this number to roughly double every 2 years for the next decade or two. Let us examine in some detail the operation of a single DRAM cell. APPLICATION EXAMPLE 6.20 SOLUTION Fig. 6.34a shows a simple model for a DRAM cell. Data are stored on the cell capacitor in true/false (or 1/0) format, where a large-capacitor voltage represents a true condition and a VI/O To sense amps VI/O + Cout Ileak 450 fF 50 pA Ccell 50 fF vcell(t) + Ileak 50 pA Ccell 50 fF - (a) vcell(t) - + 1.5 V + Cout Ccell 450 fF 50 fF - (b) 3V - (c) Figure 6.34 low voltage represents a false condition. The switch closes to allow access from the processor to the DRAM cell. Current source Ileak is an unintentional, or parasitic, current that models charge leakage from the capacitor. Another parasitic model element is the capacitance, Cout , the capacitance of the wiring connected to the output side of the cell. Both Ileak and Cout have enormous impacts on DRAM performance and design. Consider storing a true condition in the cell. A high voltage of 3.0 V is applied at node I/O and the switch is closed, causing the voltage on Ccell to quickly rise to 3.0 V. We open A simple circuit model showing (a) the DRAM memory cell, (b) the effect of charge leakage from the cell capacitor, and (c) cell conditions at the beginning of a read operation. irwin06-245-295hr.qxd 278 9-07-2010 CHAPTER 6 14:27 Page 278 C A PA C I TA N C E A N D I N D U C TA N C E the switch and the data are stored. During the store operation the charge, energy, and number of electrons, n, used are Q=CV=A50*10–15 B(3)=150 fC W = 1 CV2 = (0.5)A50 * 10-15 B A32 B = 225 fJ 2 n=Q/q=150*10–15/A1.6*10–19B=937,500 electrons Once data are written, the switch opens and the capacitor begins to discharge through Ileak. A measure of DRAM quality is the time required for the data voltage to drop by half, from 3.0 V to 1.5 V. Let us call this time tH . For the capacitor, we know vcell(t) = 1 i dt V Ccell 3 cell where, from Fig. 6.34b, icell(t)=–Ileak . Performing the integral yields vcell(t) = Ileak 1 A-I B dt = t + K Ccell 3 leak Ccell We know that at t=0, vcell=3 V. Thus, K=3 and the cell voltage is vcell(t) = 3 - Ileak tV Ccell 6.33 Substituting t=tH and vcell AtH B=1.5 V into Eq. (6.33) and solving for tH yields tH=15 ms. Thus, the cell data are gone in only a few milliseconds! The solution is rewriting the data before it can disappear. This technique, called refresh, is a must for all DRAM using this one-transistor cell. To see the effect of Cout , consider reading a fully charged Avcell=3.0 VB true condition. The I/O line is usually precharged to half the data voltage. In this example, that would be 1.5 V as seen in Fig. 6.34c. A To isolate the effect of Cout , we have removed Ileak .B Next, the switch is closed. What happens next is best viewed as a conservation of charge. Just before the switch closes, the total stored charge in the circuit is QT = Qout + Qcell = VI兾O Cout + Vcell Ccell QT = (1.5)(450 * 10-15) + (3)(50 * 10-15) = 825 fC When the switch closes, the capacitor voltages are the same (let us call it Vo) and the total charge is unchanged: and QT=825 fC=Vo Cout+Vo Ccell=Vo A450*10–15+50*10–15 B Vo = 1.65 V Thus, the change in voltage at VI兾O during the read operation is only 0.15 V. A very sensitive amplifier is required to quickly detect such a small change. In DRAMs, these amplifiers are called sense amps. How can vcell change instantaneously when the switch closes? It cannot. In an actual DRAM cell, a transistor, which has a small equivalent resistance, acts as the switch. The resulting RC time constant is very small, indicating a very fast circuit. Recall that we are not analyzing the cell’s speed—only the final voltage value, Vo . As long as the power lost in the switch is small compared to the capacitor energy, we can be comfortable in neglecting the switch resistance. By the way, if a false condition (zero volts) were read from the cell, then Vo would drop from its precharged value of 1.5 V to 1.35 V—a negative change of 0.15 V. This symmetric voltage change is the reason for precharging the I/O node to half the data voltage. Review the effects of Ileak and Cout . You will find that eliminating them would greatly simplify the refresh requirement and improve the voltage swing at node I/O when reading data. DRAM designers earn a very good living trying to do just that. irwin06-245-295hr.qxd 9-07-2010 14:27 Page 279 SECTION 6.6 DESIGN EXAMPLES 279 6.6 Design Examples We have all undoubtedly experienced a loss of electrical power in our office or our home. When this happens, even for a second, we typically find that we have to reset all of our digital alarm clocks. Let’s assume that such a clock’s internal digital hardware requires a current of 1 mA at a typical voltage level of 3.0 V, but the hardware will function properly down to 2.4 V. Under these assumptions, we wish to design a circuit that will “hold” the voltage level for a short duration, for example, 1 second. DESIGN EXAMPLE 6.21 We know that the voltage across a capacitor cannot change instantaneously, and hence its use SOLUTION appears to be viable in this situation. Thus, we model this problem using the circuit in Fig. 6.35 where the capacitor is employed to hold the voltage and the 1-mA source represents the 1-mA load. As the circuit indicates, when the power fails, the capacitor must provide all the power for the digital hardware. The load, represented by the current source, will discharge the capacitor linearly in accordance with the expression Opens on power outages 3V ± – 1-mA load C v(t) = 3.0 - 1 i(t) dt C 3 After 1 second, v(t) should be at least 2.4 V, that is, the minimum functioning voltage, and hence 2.4 = 3.0 - 1 1 (0.001) dt C 3o Solving this equation for C yields C = 1670 F Thus, from the standard capacitor values in Table 6.1, connecting three 560-F capacitors in parallel produces 1680 F. Although three 560-F capacitors in parallel will satisfy the design requirements, this solution may require more space than is available. An alternate solution involves the use of “double-layer capacitors” or what are known as Supercaps. A Web search of this topic will indicate that a company by the name of Elna America, Inc. is a major supplier of double-layer capacitors. An investigation of their product listing indicates that their DCK series of small coin-shaped supercaps is a possible alternative in this situation. In particular, the DCK3R3224 supercap is a 220-mF capacitor rated at 3.3 V with a diameter of 7 mm, or about 1/4 inch, and a thickness of 2.1 mm. Since only one of these items is required, this is a very compact solution from a space standpoint. However, there is yet another factor of importance and that is cost. To minimize cost, we may need to look for yet another alternate solution. Figure 6.35 A simple model for a power outage ridethrough circuit. irwin06-245-295hr.qxd 280 9-07-2010 CHAPTER 6 14:27 Page 280 C A PA C I TA N C E A N D I N D U C TA N C E DESIGN EXAMPLE 6.22 Let us design an op-amp circuit in which the relationship between the output voltage and two inputs is vo(t) = 5 SOLUTION 3 v1(t) dt - 2v2(t) In order to satisfy the output voltage equation, we must add two inputs, one of which must be integrated. Thus, the design equation calls for an integrator and a summer as shown in Fig. 6.36. Using the known equations for both the integrator and summer, we can express the output voltage as vo(t) = -v2(t) c R4 R4 R4 R4 1 d - c d ev1(t) dt f = v1(t) dt - c d v2(t) R3 R2 R1 C 3 R1 R2 C 3 R3 Figure 6.36 C Op-amp circuit with integrator and summer. R4 R1 + R2 + v1(t) + + v2(t) - + vo(t) R3 - - If we now compare this equation to our design requirement, we find that the following equalities must hold: R4 = 5 R1 R2 C R4 = 2 R3 Note that we have five variables and two constraint equations. Thus, we have some flexibility in our choice of components. First, we select C = 2 F, a value that is neither large nor small. If we arbitrarily select R4 = 20 k⍀, then R3 must be 10 k⍀ and furthermore R1 R2 = 2 * 109 If our third choice is R1 = 100 k⍀, then R2 = 20 k⍀. If we employ standard op-amps with supply voltages of approximately ;10 V, then all currents will be less than 1 mA, which are reasonable values. • SUMMARY ■ The important (dual) relationships for capacitors and inductors are as follows: p(t) = Cv(t) q = Cv i(t) = C v(t) = dv(t) dt p(t) = Li(t) di(t) dt 1 2 1 Cv (t) wL(t) = Li2(t) 2 2 The passive sign convention is used with capacitors and inductors. wC(t) = dv(t) dt t 1 i(x) dx C 3-q v(t) = L i(t) = di(t) ■ dt t 1 v(x) dx L 3-q ■ In dc steady state, a capacitor looks like an open circuit and an inductor looks like a short circuit. irwin06-245-295hr.qxd 9-07-2010 14:27 Page 281 281 PROBLEMS ■ The voltage across a capacitor and the current flowing through an inductor cannot change instantaneously. ■ Leakage resistance is present in practical capacitors. ■ When capacitors are interconnected, their equivalent capacitance is determined as follows: capacitors in series combine like resistors in parallel, and capacitors in parallel combine like resistors in series. • ■ When inductors are interconnected, their equivalent inductance is determined as follows: inductors in series combine like resistors in series, and inductors in parallel combine like resistors in parallel. ■ RC operational amplifier circuits can be used to differentiate or integrate an electrical signal. PROBLEMS 6.1 An uncharged 100-F capacitor is charged by a constant current of 1 mA. Find the voltage across the capacitor after 4 s. 6.9 The voltage across a 20-F capacitor is shown in Fig. P6.9. Determine the waveform for the current in the capacitor. 6.2 A 12-F capacitor has an accumulated charge of 480 C. v(t) V Determine the voltage across the capacitor. 6.3 A capacitor has an accumulated charge of 600 C with 10 5 V across it. What is the value of capacitance? 6.4 A 25-F capacitor initially charged to ⫺10 V is charged 5 by a constant current of 2.5 A. Find the voltage across the capacitor after 221 min. 6.5 The energy that is stored in a 25-F capacitor is w(t) ⫽ 12 sin2 377t. Find the current in the capacitor. 6.6 A capacitor is charged by a constant current of 2 mA and results in a voltage increase of 12 V in a 10-s interval. What is the value of the capacitance? 6.7 The current in a 100-F capacitor is shown in Fig. P6.7. Determine the waveform for the voltage across the capacitor if it is initially uncharged. 0 6 10 20 t (ms) Figure P6.9 6.10 Derive the waveform for the current in a 50-F capacitor in the voltage across the capacitor as shown in Fig. P6.10. v(t) (V) 8 i(t) (mA) 4 10 0 1 4 2 t (ms) 8 12 t (ms) Figure P6.10 Figure P6.7 6.11 If the voltage waveform across a 100-F capacitor is shown in Fig. P6.11, determine the waveform for the current. 6.8 The voltage across a 10-F capacitor is shown in Fig. P6.8. Determine the waveform for the current in the capacitor. v(t) (V) v(t) V 10 6 4 5 2 0 4 Figure P6.8 8 12 16 t (ms) 5 Figure P6.11 10 15 20 t (ms) irwin06-245-295hr.qxd 282 9-07-2010 CHAPTER 6 14:27 Page 282 C A PA C I TA N C E A N D I N D U C TA N C E 6.12 The voltage waveform across a 100-F capacitor is shown in Fig. P6.12. Derive the waveform for the current. v(t) V 6.15 The voltage across a 2-F capacitor is given by the waveform in Fig. P6.15. Find the waveform for the current in the capacitor. vC(t) (V) 12 +12 8 4 0 10 15 20 25 30 35 40 0 t (ms) 10 20 30 40 50 t (s) –12 Figure P6.12 Figure P6.15 6.13 The current flowing through a 5-F capacitor is shown in Fig. P6.13. Find the energy stored in the capacitor at t=1.4 ms, t=3.3 ms, t=4.3 ms, t=6.7 ms, and t=8.5 ms. i(t) (mA) 6.16 The voltage across a 2-F capacitor is given by the waveform in Fig. P6.16. Compute the current waveform. v(t) (V) 2 15 3 6 t (ms) –12 10 Figure P6.16 5 4 6 6.17 Draw the waveform for the current in a 24-F capacitor 7 t (ms) 1 2 3 5 when the capacitor voltage is as described in Fig. P6.17. 8 v(t) (V) –5 6 + vc(t) i(t) 100 5 F 0 - 160 t (s) 60 –4 Figure P6.13 6.14 The voltage across a 25-F capacitor is shown in Figure P6.17 Fig. P6.14. Determine the current waveform. 6.18 The voltage across a 10-F capacitor is given by the v(t) (V) waveform in Fig. P6.18. Plot the waveform for the capacitor current. 20 0.8 0 –20 Figure P6.14 0.2 0.4 0.6 1.0 1.2 t (ms) v(t) (V) 12 5 –12 Figure P6.18 10 t (ms) irwin06-245-295hr.qxd 9-07-2010 14:27 Page 283 PROBLEMS 6.19 The waveform for the current in a 50-F capacitor is 283 6.20 The waveform for the current in a 50-F initially shown in Fig. P6.19. Determine the waveform for the capacitor voltage. uncharged capacitor is shown in Fig. P6.20. Determine the waveform for the capacitor’s voltage. i(t) (mA) 10 i(t) (mA) 10 0 0 0 10 20 30 40 10 20 30 40 50 t (ms) t (ms) –10 Figure P6.19 Figure P6.20 6.21 The waveform for the voltage across 100-F capacitor shown Fig. 6.21a is given in Fig. 6.21b. Determine the following quantities: (a) the energy stored in the capacitor at t = 2.5 ms, (b) the energy stored in the capacitor at t = 5.5 ms, (c) ic(t) at t = 1.5 ms, (d) ic(t) at t = 4.75 ms, and (e) ic(t) at t = 7.5 ms. v(t) (V) 15 5 4 v(t) ± – 100 µF 5 t (ms) ic(t) 1 2 3 6 7 8 –5 (b) (a) Figure P6.21 6.22 The current in an inductor changed from 0 to 200 mA in 4 ms and induces a voltage of 100 mV. What is the value of the inductor? 6.23 The current in a 100-mH inductor is i(t) = 2 sin 377t A. Find (a) the voltage across the inductor and (b) the expression for the energy stored in the element. 6.27 The voltage across a 2-H inductor is given by the waveform shown in Fig. P6.27. Find the waveform for the current in the inductor. v(t) (V) 5 6.24 If the current i(t) = 1.5t A flows through a 2-H 0 inductor, find the energy stored at t = 2s. –5 6.25 The current in a 25-mH inductor is given by the Figure P6.27 expressions i(t) = 0 t 6 0 i(t) = 10(1 - e - t) mA t 7 0 Find (a) the voltage across the inductor and (b) the expression for the energy stored in it. 6.26 Given the data in the previous problem, find the 2 4 6 t (s) 6.28 The voltage across a 4-H inductor is given by the waveform shown in Fig. P6.28. Find the waveform for the current in the inductor. v(t) = 0, t 6 0. v(t) (mV) 2.4 voltage across the inductor and the energy stored in it after 1 s. 0 Figure P6.28 10 20 30 40 50 t (ms) irwin06-245-295hr.qxd 284 9-07-2010 CHAPTER 6 14:27 Page 284 C A PA C I TA N C E A N D I N D U C TA N C E 6.29 The current in a 30-mH inductor is shown in Fig. P6.29. 6.32 The current in a 200-mH inductor is shown in Derive the waveform for the inductor voltage. Fig. P6.32. Determine the waveform for the inductor voltage. i(t) (mA) i(t) (mA) 60 120 40 60 t (ms) 10 20 30 40 50 60 20 70 2 4 6 8 t (ms) 10 Figure P6.29 12 14 16 –20 –40 6.30 The current waveform in a 40-mH inductor is shown in Fig. P6.30. Derive the waveform for the inductor voltage. Figure P6.32 6.33 The waveform for the current flowing through a 0.5-H inductor is shown in the plot in Fig. P6.33. Accurately sketch the inductor voltage versus time. Determine the following quantities: (a) the energy stored in the inductor at t=1.7 ms, (b) the energy stored in the inductor at t=4.2 ms, and (c) the power absorbed by the inductor at t=1.2 ms, t=2.8 ms, and t=5.3 ms. i(t) (mA) 20 15 i(t) (mA) 10 10 5 5 4 t (ms) t (ms) 2 4 6 8 10 1 12 2 3 5 6 –5 Figure P6.30 –10 6.31 If the current in a 60-mH inductor is given by the wave- + form in Fig. P6.31, compute the waveform for the inductor voltage. vL(t) i(t) 0.5 H - i(t) (mA) Figure P6.33 8 6.34 The current in a 10-mH inductor is shown in Fig. P6.34. 6 Determine the waveform for the voltage across the inductor. i(t) (mA) 4 0 1 2 3 4 5 6 2 t (ms) t (ms) 2 Figure P6.31 4 6 8 10 –12 Figure P6.34 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 285 PROBLEMS 6.35 The current in a 50-mH inductor is given in Fig. P6.35. 285 6.39 Find the possible capacitance range of the following Sketch the inductor voltage. capacitors. i(t) (mA) (a) 0.0068 F with a tolerance of 10%. (b) 120 pF with a tolerance of 20%. 100 (c) 39 F with a tolerance of 20%. 4 0 2 6 8 10 t (ms) –100 Figure P6.35 6.40 Find the possible inductance range of the following 6.36 The current in a 50-mH inductor is shown in Fig. P6.36. Find the voltage across the inductor. inductors. (a) 10 mH with a tolerance of 10%. (b) 2.0 nH with a tolerance of 5%. i(t) (mA) (c) 68 H with a tolerance of 10%. +10 0 20 40 60 70 80 t (ms) –20 Figure P6.36 6.41 The capacitor in Fig. P6.41a is 51 nF with a tolerance of 6.37 Draw the waveform for the voltage across a 24-mH 10%. Given the voltage waveform in Fig. 6.41b, graph the current i(t) for the minimum and maximum capacitor values. inductor when the inductor current is given by the waveform shown in Fig. P6.37. i(t) i(t) (A) 8 v(t) ± – 4 0.6 0.9 0.3 –2 1.1 C t (s) (a) Figure P6.37 60 40 in Fig. P6.38. Plot the voltage across the inductor. i(t) (mA) v(t) (V) 6.38 The current in a 4-mH inductor is given by the waveform 20 0 –20 0.12 –40 –60 0 0.5 1.0 1 t (ms) 2 3 Time (ms) (b) Figure P6.38 4 Figure P6.41 5 6 7 irwin06-245-295hr.qxd 286 9-07-2010 CHAPTER 6 14:27 Page 286 C A PA C I TA N C E A N D I N D U C TA N C E 6.42 The inductor in Fig. P6.42a is 4.7 H with a tolerance 6.45 Given the network in Fig. P6.45, find the power dissi- of 20%. Given the current waveform in Fig. 6.42b, graph the voltage v(t) for the minimum and maximum inductor values. pated in the 3-⍀ resistor and the energy stored in the capacitor. 3⍀ 2H ± v(t) L i(t) 12 V 3H 4⍀ ± – 6⍀ 6A 2F – (a) Figure P6.45 15 10 6.46 Calculate the energy stored in the inductor in the circuit shown in Fig. P6.46. i(t) (mA) 5 0 1 k⍀ 5 k⍀ 4 k⍀ –5 –10 + – –15 0 10 20 30 40 50 60 70 80 200 mH 20 V 6 k⍀ 7 kV Time (ms) (b) Figure P6.46 Figure P6.42 6.47 Calculate the energy stored in both the inductor and the capacitor shown in Fig. P6.47. 6.43 If the total energy stored in the circuit in Fig. P6.43 is 10 ⍀ 80 mJ, what is the value of L? L 25 V 200 ⍀ 1A 80 F + – 15 ⍀ 15 ⍀ 0.005 F 0.5 H 50 ⍀ Figure P6.47 6.48 Given a 1-, 3-, and 4-F capacitor, can they be intercon- Figure P6.43 nected to obtain an equivalent 2-F capacitor? 6.49 Find the total capacitance CT of the network in Fig. P6.49. 6.44 Find the value of C if the energy stored in the capacitor in Fig. P6.44 equals the energy stored in the inductor. 4 F C CT 12 F 100 ⍀ 12 V ± – Figure P6.44 200 ⍀ 1 F 0.1 H 2 F Figure P6.49 3 F irwin06-245-295hr.qxd 9-07-2010 14:27 Page 287 PROBLEMS 6.50 Find the total capacitance CT of the network in 287 6.54 Find CT in the network in Fig. P6.54. Fig. P6.50. 12 F 3 F 3 F CT 6 F 4 F 3 F 3 F 3 F 6 F 4 F CT Figure P6.50 6 F 6.51 Find CT in the network shown in Fig. P6.51. Figure P6.54 3 F 6 F 6 F 8 F 6.55 Determine CT in the circuit in Fig. P6.55 if all capacitors in the network are 6 F. 4 F 12 F CT Figure P6.51 6.52 Find CT in the circuit in Fig. P6.52. CT 4 F 5 F CT 1 F 5 F Figure P6.55 16 F 3 F 6.56 Find CT in the circuit in Fig. P6.56 if all capacitors Figure P6.52 are 6 F. 6.53 Determine the value of CT in the circuit in Fig. P6.53. 12 F 9 F CT 6 F CT 11 F 6 F 5 F 3 F Figure P6.53 Figure P6.56 irwin06-245-295hr.qxd 288 9-07-2010 CHAPTER 6 14:27 Page 288 C A PA C I TA N C E A N D I N D U C TA N C E 6.57 Determine the value of CT in the circuit in Fig. P6.57 if all capacitors are 12 F. 6.61 If Ceq = 4 F in the circuit in Fig. P6.61, calculate C. 6 F Ceq 7 F C CT Figure P6.61 Figure P6.57 6.58 If the total capacitance of the network in Fig. P6.58 is 10F, find the value of CT. 6.62 Find the total capacitance CT shown in the network in Fig. P6.62. 12 F 12 F 6 F CT ⫽ 10 F 6 F C 4 F 8 F 4 F 2 F Figure P6.58 6.59 In the network in Fig. 6.59, if CT = 4 F, find the 4 F 2 F CT value of C. Figure P6.62 4 F CT ⫽ 4 F 2 F C 12 F 6.63 In the network in Fig. P6.63, find the capacitance CT if Figure P6.59 (a) the switch is open and (b) the switch is closed. 6.60 Find the value of C in Fig. 6.60. 12 F 6 F 3 F 12 F 12 F CT 2 F 10 F CT ⫽ 4 F 6 F 6 F 3 F C Figure P6.60 Figure P6.63 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 289 PROBLEMS 289 6.67 The two capacitors shown in Fig. P6.67 have been 6.64 Select the value of C to produce the desired total capacitance of CT = 10 F in the circuit in Fig. P6.64. connected for some time and have reached their present values. Find V0. + Vo 12 F + 16 V - 4 F C - CT=10 F 8 F 16 F Figure P6.67 Figure P6.64 6.65 Select the value of C to produce the desired total capacitance of CT = 1 F in the circuit in Fig. P6.65. C 6.68 The three capacitors shown in Fig. P6.68 have been connected for some time and have reached their present values. Find V1 and V2. 1 F C + CT V1 - 1 F 2 F 8 F + 12 V - 1 F + V2 4 F - Figure P6.65 6.66 The two capacitors in Fig. P6.66 were charged and then Figure P6.68 connected as shown. Determine the equivalent capacitance, the initial voltage at the terminals, and the total energy stored in the network. 6.69 Determine the inductance at terminals A-B in the network in Fig. P6.69. 6V + 12 F 1 mH A 4 mH + 2V - 2 mH 4 F 3 mH B Figure P6.66 12 mH 2 mH Figure P6.69 4 mH irwin06-245-295hr.qxd 290 9-07-2010 CHAPTER 6 14:27 Page 290 C A PA C I TA N C E A N D I N D U C TA N C E 6.70 Find the total inductance at the terminals of the network 6.73 Find LT in the circuit in Fig. P6.73. All inductors are 12 H. in Fig. P6.70. 12 mH 12 mH 6 mH LT 4 mH 2 mH LT Figure P6.70 Figure P6.73 6.71 Find LT in the circuit in Fig. P6.71. 6.74 Find LT in the circuit in Fig. P6.74. LT 12 H 4 H 6 H 5 H 3 H 2 H 4 H 5 H 4 H 8 H LT 6 H 12 H 1 H 7 H Figure P6.71 6 H 2 H Figure P6.74 6.72 Find LT in the circuit in Fig. P6.72. 12 H 6.75 Find LT in the circuit in Fig. P6.75. 2 H 4 H 5 H LT 2 H 8 H 4 H 6 H 4 H 6 H 3 H LT 6 H 4 H 3 H Figure P6.72 9 H Figure P6.75 3 H 12 H 12 H irwin06-245-295hr.qxd 9-07-2010 14:27 Page 291 291 PROBLEMS 6.76 If the total inductance, LT, in the network in Fig. P6.76 is 5 H, find the value of L. 2 H equivalent inductance at terminals A-B with terminals C-D short circuited, and (b) the equivalent inductance at terminals C-D with terminals A-B open circuited. 1 H 20 mH 3 H LT ⫽ 5 H A 10 H C 3 H 6 H 1 H 6.79 Given the network shown in Fig. P6.79, find (a) the 5 mH 12 mH L B Figure P6.76 D 6 mH Figure P6.79 6.77 If the total inductance, LT, of the network in Fig. P6.77 is 6 H, find the value of L. 2 H 6.80 Find the value of L in the network in Fig. P6.80 so that the total inductance, LT, will be 2 mH. 8 H 4 mH 1 H LT ⫽ 4 H 3 H 4 H LT 2 mH L L 6 mH 4 H 18 H 9 H Figure P6.80 Figure P6.77 6.81 A 20-mH inductor and a 12-mH inductor are connected in series with a 1-A current source. Find (a) the equivalent inductance and (b) the total energy stored. 6.78 Find LT in the network in Fig. P6.78 (a) with the switch open and (b) with the switch closed. All inductors are 12 mH. 6.82 If the capacitors shown in Fig. P6.82 have been connected for some time and have reached. Their present values, determine (a) the voltage V0 and (b) the total energy stored in the capacitors. LT 12 V +3 F 6 F Figure P6.78 Figure P6.82 Vo irwin06-245-295hr.qxd 292 9-07-2010 CHAPTER 6 14:27 Page 292 C A PA C I TA N C E A N D I N D U C TA N C E 6.83 If the capacitors in the circuit in Fig. P6.83 have been connected for some time and have reached their present values, calculate (a) the voltage V1 and (b) the total energy stored in the capacitors. 6 F 12 F 6.86 If the capacitors in the circuit in Fig. P6.86 have been connected for some time and have reached their present values, (a) calculate the voltages V1 and V2 and (b) determine the total energy stored in the capacitors. + + V1 + 6V - 10 F 9 F 36 V 3 F - + V -1 + V -2 Figure P6.83 Figure P6.86 6.87 For the network in Fig. P6.87 vS1(t) = 80 cos 377t V and vS2(t) = 40 cos 377t V. Find v0(t) . 1 F 6.84 If the capacitors shown in Fig. P6.84 have been connected for some time and the voltage has reached its present value, determine (a) the voltages V1 and V2 and (b) the total energy stored in the capacitors. 20 k⍀ 10 k⍀ + ± vS1(t) – + ± – vS2(t) vo(t) - + + 24 V + - V1 12 F V2 4 F - Figure P6.87 6.88 If the input to the network shown in Fig. P6.88a is given by the waveform in Fig. P.6.88b, determine the output waveform v0(t) if v0(0)=0. Figure P6.84 + - + + 40 k⍀ vi(t) - 6.85 If the capacitors shown in Fig. P6.85 have been connected for some time and the voltage has reached its present value, find (a) the voltages V1 and V2 and (b) the total energy stored in the capacitors. vo(t) 100 k⍀ - (a) vi (t) 2 1 + + 18 V + - V1 V2 - 2 3 F 6 F 3 4 1 –2 Figure P6.88 6 7 t (s) –1 (b) Figure P6.85 5 irwin06-245-295hr.qxd 9-07-2010 14:27 Page 293 293 PROBLEMS 6.89 The input to the network shown in Fig. P6.89a is shown 6.91 Sketch the output voltage of the network in Fig. P6.91a if the input is given by the waveform in Fig. 6.91b. in Fig. P6.89b. Derive the waveform for the output voltage v0(t) if v0(0)=0. 2 k⍀ 200 F 1 µF + + 20 k⍀ + + + vi(t) vo(t) - - + vi(t) vo(t) - (a) (a) vi (t) vi (t) (V) 3 3 2 1 3 1 4 2 2 t (s) 4 6 –1 8 t (b) Figure P6.91 (b) Figure P6.89 6.92 Sketch the output voltage of the network in Fig. P6.92a if the input is given by the waveform in Fig. P6.92b. 4 k⍀ 2 µF 6.90 Show that the circuit shown in Fig. P6.90 acts like a differentiator with an output voltage of vo(t) = -RC Assume an ideal op-amp. + + dv0(t) dt + vi(t) vo(t) - (a) C + vi (t) (V) R + 3 + 2 vi(t) vo(t) 1 - 1 2 3 4 (b) Figure P6.90 Figure P6.92 5 6 7 t (ms) irwin06-245-295hr.qxd 294 9-07-2010 CHAPTER 6 14:27 Page 294 C A PA C I TA N C E A N D I N D U C TA N C E 6.93 Given the network in Fig. P6.93a, 6.94 An integrator is required that has the following performance: (a) Determine the equation for the closed-loop gain |G| = 2 v0 (t) = 106 v0 2. vi (b) Sketch the magnitude of the closed-loop gain as a function of frequency if R1=1 k⍀, R2=10 k⍀, and C=2 F. 3 vs dt where the capacitor values must be greater than 10 nF and the resistor values must be greater than 10 k⍀. (a) Design the integrator. (b) If ⫾10-V supplies are used, what are the maximum C and minimum values of v0? (c) Suppose vs ⫽ 1 V. What is the rate of change of v0? R2 R1 + + + vi(t) vo(t) - - Figure P6.93 TYPICAL PROBLEMS FOUND ON THE FE EXAM • 6PFE-1 Given three capacitors with values 2-F, 4-F, and 6PFE-2 The current pulse shown in Fig. 6PFE-2 is applied 6-F, can the capacitors be interconnected so that the combination is an equivalent 3-F capacitor? to a 1-F capacitor. What is the energy stored in the electric field of the capacitor? a. Yes. The capacitors should be connected as shown. 0 J, t ⱕ 0 a. w(t) = • 10 * 106t2J, 0 6 t ⱕ 1 s 10 J, t 7 1 s 6 µF 2 µF 4 µF Ceq 0 J, t ⱕ 0 b. w(t) = • 6 * 106t J, 0 6 t ⱕ 1 s 6 J, t 7 1 s 0 J, t ⱕ 0 c. w(t) = • 18 * 106t2J, 0 6 t ⱕ 1 s 18 J, t 7 1 s b. Yes. The capacitors should be connected as shown. 0 J, t ⱕ 0 d. w(t) = • 30 * 106t J, 0 6 t ⱕ 1 s 30 J, t 7 1 s 4 µF 2 µF 6 µF Ceq i(t) (A) c. Yes. The capacitors should be connected as shown. 6 2 µF 4 µF 6 µF Ceq 0 Figure 6PFE-2 d. No. An equivalent capacitance of 3 F is not possible with the given capacitors. 1 t (s) irwin06-245-295hr.qxd 9-07-2010 14:27 Page 295 TYPICAL PROBLEMS FOUND ON THE FE EXAM 6PFE-3 The two capacitors shown in Fig. 6FE-3 have been connected for some time and have reached their present values. Determine the unknown capacitor Cx. a. 20 F + b. 30 F + 8V - c. 10 F d. 90 F 60 F 24 V Figure 6PFE-3 Cx - 6PFE-4 What is the equivalent inductance of the network in Fig. 6PFE-4? a. 9.5 mH b. 2.5 mH c. 6.5 mH d. 3.5 mH 2 mH 3 mH 6PFE-5 The current source in the circuit in Fig. 6PFE-5 has the following operating characteristics: i(t) = e 0 A, t 6 0 20te-2t A, t 7 0 What is the voltage across the 10-mH inductor expressed as a function of time? a. v(t) = e 0 V, t 6 0 0.2e-2t - 4te-2t V, t 7 0 b. v(t) = e 0 V, t 6 0 2e-2t + 4te-2t V, t 7 0 c. v(t) = e 0 V, t 6 0 -0.2te-2t + 0.4e-2t V, t 7 0 d. v(t) = e 0 V, t 6 0 -2te-2t V, t 7 0 12 mH 3 mH 6 mH + 9 mH Leq Figure 6PFE-4 295 i(t) Figure 6PFE-5 L v(t) - irwin07_296-368hr.qxd 28-07-2010 11:34 Page 296 7 C H A PT E R FIRST- AND SECOND-ORDER TRANSIENT CIRCUITS THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Be able to calculate initial values for inductor cur- rents and capacitor voltages in transient circuits ■ Know how to calculate voltages and currents in first- order transient circuits ■ Know how to calculate voltages and currents in second-order transient circuits E.S.A./SIPA/NewsCom S Sea Ice Measurements Data clearly demonstrate an ongoing using radar backscatter to map transitions in sea ice thickness. rise in global temperature. NASA’s Goddard Institute for Space Sea ice growth and melt measurements from radar during all Studies has reported that we have just experienced the warmest summer-to-winter seasons confirm that the ice is disappearing. decade (2000 to 2009) on record. Measuring the environmental This chapter describes voltage and current transitions in cir- impact of this change – and deciding how to respond to it – is a cuits resulting from switching of constant-level sources from one major challenge facing scientists and policymakers today and value to another. Capacitors, inductors, and resistors are con- will be for years to come. nected in series or parallel, modeling transitions for simple to Electronic devices are essential to efforts to measure impacts more complex circuits. Extending to a square-wave pulse voltage such as the reduction in polar ice. Polar ice regulates global cli- source demonstrates multiple switching dynamics. The transi- mate because it reflects about 80% of the incoming sunlight. tions in circuit voltages and currents that you will study in this Radar altimetry data from satellites is being used to measure the chapter are principles that you can see applied in a wide array of effect of warming on polar glaciers. Low-flying aircraft making devices. While some seem mundane, others, such as those that repeated runs across Arctic, Antarctic, and Greenland ice beds are measure sea ice, have a major impact on our future. 296 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 297 SECTION 7.1 In this chapter we perform what is normally referred to as a transient analysis. We begin our analysis with first-order circuits—that is, those that contain only a single storage element. When only a single storage element is present in the network, the network can be described by a first-order differential equation. Our analysis involves an examination and description of the behavior of a circuit as a function of time after a sudden change in the network occurs due to switches opening or closing. Because of the presence of one or more storage elements, the circuit response to a sudden change will go through a transition period prior to settling down to a steady-state value. It is this transition period that we will examine carefully in our transient analysis. One of the important parameters that we will examine in our transient analysis is the circuit’s time constant. This is a very important network parameter because it tells us how fast the circuit will respond to changes. We can contrast two very different systems to obtain a feel for the parameter. For example, consider the model for a room air-conditioning system and the model for a single-transistor stage of amplification in a computer chip. If we change the setting for the air conditioner from 70 degrees to 60 degrees, the unit will come on and the room will begin to cool. However, the temperature measured by a thermometer in the room will fall very slowly and, thus, the time required to reach the desired temperature is long. However, if we send a trigger signal to a transistor to change state, the action may take only a few nanoseconds. These two systems will have vastly different time constants. Our analysis of first-order circuits begins with the presentation of two techniques for performing a transient analysis: the differential equation approach, in which a differential equation is written and solved for each network, and a step-by-step approach, which takes advantage of the known form of the solution in every case. In the second-order case, both an inductor and a capacitor are present simultaneously, and the network is described by a second-order differential equation. Although the RLC circuits are more complicated than the first-order single storage circuits, we will follow a development similar to that used in the first-order case. Our presentation will deal only with very simple circuits, since the analysis can quickly become complicated for networks that contain more than one loop or one nonreference node. Furthermore, we will demonstrate a much simpler method for handling these circuits when we cover the Laplace transform later in this book. We will analyze several networks in which the parameters have been chosen to illustrate the different types of circuit response. Finally, a number of application-oriented examples are presented and discussed. We begin our discussion by recalling that in Chapter 6 we found that capacitors and inductors were capable of storing electric energy. In the case of a charged capacitor, the energy is stored in the electric field that exists between the positively and negatively charged plates. This stored energy can be released if a circuit is somehow connected across the capacitor that provides a path through which the negative charges move to the positive charges. As we know, this movement of charge constitutes a current. The rate at which the energy is discharged is a direct function of the parameters in the circuit that is connected across the capacitor’s plates. As an example, consider the flash circuit in a camera. Recall that the operation of the flash circuit, from a user standpoint, involves depressing the push button on the camera that triggers both the shutter and the flash and then waiting a few seconds before repeating the process to take the next picture. This operation can be modeled using the circuit in Fig. 7.1a. The voltage source and resistor RS model the batteries that power the camera and flash. The capacitor models the energy storage, the switch models the push button, and finally the resistor R models the xenon flash lamp. Thus, if the capacitor is charged, when the switch is closed, the capacitor voltage drops and energy is released through the xenon lamp, producing the flash. In practice this energy release takes about a millisecond, and the discharge time is a function of the elements in the circuit. When the push button is released and the switch is then opened, the battery begins to recharge the capacitor. Once again, the time required to charge the capacitor is a function of the circuit elements. The discharge and charge cycles are graphically illustrated in Fig. 7.1b. Although the discharge time is very fast, it is not instantaneous. To provide further insight into this phenomenon, consider what INTRODUCTION 297 7.1 Introduction irwin07_296-368hr.qxd 298 28-07-2010 CHAPTER 7 11:34 Page 298 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Figure 7.1 Discharge time vC(t) Diagrams used to describe a camera’s flash circuit. Vo RS + VS Charge time C vC(t) R - t [(R) Xenon lamp] (a) (b) + vC(t) - C R (c) we might call a free-body diagram of the right half of the network in Fig. 7.1a, as shown in Fig. 7.1c (that is, a charged capacitor that is discharged through a resistor). When the switch is closed, KCL for the circuit is C dvC(t) vC(t) + dt R = 0 or dvC(t) + dt 1 v (t) = 0 RC C In the next section we will demonstrate that the solution of this equation is vC(t) = Vo e-t兾RC Note that this function is a decaying exponential and the rate at which it decays is a function of the values of R and C. The product RC is a very important parameter, and we will give it a special name in the following discussions. 7.2 First-Order Circuits G E N E R A L F O R M O F T H E R E S P O N S E EQ U AT I O N S In our study of first-order transient circuits we will show that the solution of these circuits (i.e., finding a voltage or current) requires us to solve a first-order differential equation of the form dx(t) dt + ax(t) = f(t) 7.1 Although a number of techniques may be used for solving an equation of this type, we will obtain a general solution that we will then employ in two different approaches to transient analysis. A fundamental theorem of differential equations states that if x(t) = xp(t) is any solution to Eq. (7.1), and x(t) = xc(t) is any solution to the homogeneous equation dx(t) dt + ax(t) = 0 7.2 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 299 SECTION 7.2 F I R S T- O R D E R C I R C U I T S then x(t) = xp(t) + xc(t) 7.3 is a solution to the original Eq. (7.1). The term xp(t) is called the particular integral solution, or forced response, and xc(t) is called the complementary solution, or natural response. At the present time we confine ourselves to the situation in which f(t)=A (i.e., some constant). The general solution of the differential equation then consists of two parts that are obtained by solving the two equations dx p(t) dt dx c(t) dt + ax p(t) = A 7.4 + ax c(t) = 0 7.5 Since the right-hand side of Eq. (7.4) is a constant, it is reasonable to assume that the solution xp(t) must also be a constant. Therefore, we assume that xp(t) = K1 7.6 Substituting this constant into Eq. (7.4) yields K1 = A a 7.7 = -a 7.8 Examining Eq. (7.5), we note that dx c(t)兾dt xc(t) This equation is equivalent to d C ln xc(t)D = -a dt Hence, ln xc(t) = -at + c and therefore, xc(t) = K2 e-at 7.9 Thus, a solution of Eq. (7.1) is x(t) = xp(t) + xc(t) = A + K2 e-at a 7.10 The constant K2 can be found if the value of the independent variable x(t) is known at one instant of time. Eq. (7.10) can be expressed in general in the form x(t) = K1 + K2 e-t兾 7.11 Once the solution in Eq. (7.11) is obtained, certain elements of the equation are given names that are commonly employed in electrical engineering. For example, the term K1 is referred to as the steady-state solution: the value of the variable x(t) as t S q when the second term becomes negligible. The constant is called the time constant of the circuit. Note that the second term in Eq. (7.11) is a decaying exponential that has a value, if 7 0, of K2 for t=0 and a value of 0 for t=q. The rate at which this exponential decays is determined by the time constant . A graphical picture of this effect is shown in Fig. 7.2a. As can be seen from the figure, the value of xc(t) has fallen from K2 to a value of 0.368K2 in one time constant, a drop of 63.2%. In two time constants the value of xc(t) has fallen to 0.135K2 , 299 irwin07_296-368hr.qxd 300 28-07-2010 CHAPTER 7 11:34 Page 300 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Figure 7.2 xc(t)=K2e–t/ Time-constant illustrations. K2 0.368K2 0.632 GG 0 2 3 4 3 4 t (a) e–t/ 1.0 0.8 0.6 0.4 0.2 =4 s =0.5 s 0 1 2 t (b) a drop of 63.2% from the value at time t = . This means that the gap between a point on the curve and the final value of the curve is closed by 63.2% each time constant. Finally, after five time constants, xc(t) = 0.0067K2 , which is less than 1%. An interesting property of the exponential function shown in Fig. 7.2a is that the initial slope of the curve intersects the time axis at a value of t = . In fact, we can take any point on the curve, not just the initial value, and find the time constant by finding the time required to close the gap by 63.2%. Finally, the difference between a small time constant (i.e., fast response) and a large time constant (i.e., slow response) is shown in Fig. 7.2b. These curves indicate that if the circuit has a small time constant, it settles down quickly to a steady-state value. Conversely, if the time constant is large, more time is required for the circuit to settle down or reach steady state. In any case, note that the circuit response essentially reaches steady state within five time constants (i.e., 5). Note that the previous discussion has been very general in that no particular form of the circuit has been assumed—except that it results in a first-order differential equation. A N A LYS I S T EC H N I Q U E S The Differential Equation Approach Eq. (7.11) defines the general form of the solution of first-order transient circuits; that is, it represents the solution of the differential equation that describes an unknown current or voltage anywhere in the network. One of the ways that we can arrive at this solution is to solve the equations that describe the network behavior using what is often called the state-variable approach. In this technique we write the equation for the voltage across the capacitor and/or the equation for the current through the inductor. Recall from Chapter 6 that these quantities cannot change instantaneously. Let us first illustrate this technique in the general sense and then examine two specific examples. Consider the circuit shown in Fig. 7.3a. At time t = 0, the switch closes. The KCL equation that describes the capacitor voltage for time t 7 0 is C v(t) - VS dv(t) dt + R = 0 or v(t) dv(t) dt + RC = VS RC irwin07_296-368hr.qxd 28-07-2010 11:34 Page 301 SECTION 7.2 +vR(t)- t=0 R VS ± – 301 v(t), vR(t) v(t) t=0 F I R S T- O R D E R C I R C U I T S VS R VS ± – C L i(t) t (a) (b) From our previous development, we assume that the solution of this first-order differential equation is of the form v(t) = K1 + K2 e-t兾 Substituting this solution into the differential equation yields - K2 -t兾 K2 -t兾 K1 VS + e + = e RC RC RC Equating the constant and exponential terms, we obtain K1 = VS = RC Therefore, v(t) = VS + K2 e-t兾RC where VS is the steady-state value and RC is the network’s time constant. K2 is determined by the initial condition of the capacitor. For example, if the capacitor is initially uncharged (that is, the voltage across the capacitor is zero at t = 0), then 0 = VS + K2 or K2 = -VS Hence, the complete solution for the voltage v(t) is v(t) = VS - VS e-t兾RC The circuit in Fig. 7.3b can be examined in a similar manner. The KVL equation that describes the inductor current for t 7 0 is L di(t) + Ri(t) = VS dt A development identical to that just used yields i(t) = R VS -a b t + K2 e L R where VS兾R is the steady-state value and L/R is the circuit’s time constant. If there is no initial current in the inductor, then at t = 0 0 = VS + K2 R and K2 = -VS R (c) Figure 7.3 (a) RC circuit, (b) RL circuit, (c) plot of the capacitor voltage in (a) and resistor voltage in (b). irwin07_296-368hr.qxd 302 28-07-2010 CHAPTER 7 11:34 Page 302 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Hence, VS - RL t VS e R R i(t) = is the complete solution. Note that if we wish to calculate the voltage across the resistor, then vR(t) = Ri (t) = VS A 1 - e R -Lt B Therefore, we find that the voltage across the capacitor in the RC circuit and the voltage across the resistor in the RL circuit have the same general form. A plot of these functions is shown in Fig. 7.3c. EXAMPLE Consider the circuit shown in Fig. 7.4a. Assuming that the switch has been in position 1 for a long time, at time t=0 the switch is moved to position 2. We wish to calculate the current i(t) for t 7 0. 7.1 t=0 1 v(t) R1 i(t) 6 k 6 k 12 V ± V – S 2 C 100 F R2 3 k 12 V ± – vC(0–) 12 V 3 k - (b) t=0– (a) R1 + v(t) i(t) (mA) i(t) 4 — 3 ± – C R2 t (c) (d) Figure 7.4 Analysis of RC circuits. SOLUTION At t = 0- the capacitor is fully charged and conducts no current since the capacitor acts like an open circuit to dc. The initial voltage across the capacitor can be found using voltage division. As shown in Fig. 7.4b, vC(0-) = 12 a 3k b = 4V 6k + 3k The network for t 7 0 is shown in Fig. 7.4c. The KCL equation for the voltage across the capacitor is v(t) R1 + C v(t) dv(t) dt + R2 Using the component values, the equation becomes dv(t) dt + 5v(t) = 0 = 0 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 303 SECTION 7.2 F I R S T- O R D E R C I R C U I T S 303 The form of the solution to this homogeneous equation is v(t) = K2 e-t兾 If we substitute this solution into the differential equation, we find that = 0.2 s. Thus, v(t) = K2 e-t兾0.2 V Using the initial condition vC(0-) = vC(0+) = 4 V, we find that the complete solution is v(t) = 4e-t兾0.2 V Then i(t) is simply i(t) = v(t) R2 or i(t) = 4 -t兾0.2 e mA 3 The switch in the network in Fig. 7.5a opens at t=0. Let us find the output voltage vo(t) for t 7 0. At t = 0- the circuit is in steady state and the inductor acts like a short circuit. The initial current through the inductor can be found in many ways; however, we will form a Thévenin equivalent for the part of the network to the left of the inductor, as shown in Fig. 7.5b. From this network we find that I1 = 4 A and Voc = 4 V. In addition, RTh=1 ⍀. Hence, iL(0-) obtained from Fig. 7.5c is iL(0-) = 4兾3 A. The network for t 7 0 is shown in Fig. 7.5d. Note that the 4-V independent source and the 2-ohm resistor in series with it no longer have any impact on the resulting circuit. The KVL equation for the circuit is -VS1 + R1 i(t) + L di(t) dt + R3 i(t) = 0 which with the component values reduces to di(t) dt + 2i(t) = 6 The solution to this equation is of the form i(t) = K1 + K2 e-t兾 which when substituted into the differential equation yields K1 = 3 = 1兾2 Therefore, i(t) = A3 + K2 e-2t B A Evaluating this function at the initial condition, which is iL(0-) = iL(0+) = i(0) = 4兾3 A we find that K2 = -5 3 EXAMPLE 7.2 SOLUTION irwin07_296-368hr.qxd 304 28-07-2010 CHAPTER 7 11:34 Page 304 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S R1 L 2 2H + I1 t=0 ± VS 1 – 12 V R2 R3 2 2 vo(t) 12 V 4V + 2 – ± V S2 2 ± – Voc – ± - 4V - (a) (b) iL(0–) + 1 4V ± – 2 vo(0–) (c) i(t) L R1 2H 2 12 V ± VS 1 – 2 + R3 2 vo(t) vo(t) (V) 6 4V – ± 8 — 3 (d) Figure 7.5 (e) Hence, Analysis of an RL circuit. t i(t) = a 3 - 5 -2t e bA 3 vo(t) = 6 - 10 -2t e V 3 and then A plot of the voltage vo(t) is shown in Fig. 7.5e. irwin07_296-368hr.qxd 28-07-2010 11:34 Page 305 SECTION 7.2 F I R S T- O R D E R C I R C U I T S Learning Assessments E7.1 Find vC(t) for t 7 0 in the circuit shown in Fig. E7.1. ANSWER: vC(t) = 8e-t兾0.6 V. t=0 3 k 4 k + 12 V + ± – vc(t) 100 F - 2 k vo(t) - Figure E7.1 E7.2 Use the differential equation approach to find vo(t) for t>0 in Fig. E7.2. Plot the response. t=0 6 k 9V + – + vo(t) 6 k - 2.5 F 6 k ANSWER: vo(t)= 12-5e-t/0.015 V. 2 mA Figure E7.2 E7.3 In the circuit shown in Fig. E7.3, the switch opens at t = 0. Find i1(t) for t 7 0. t=0 12 V ANSWER: i1(t) = 1e-9t A. i1(t) 6 12 i2(t) 2H ± – Figure E7.3 E7.4 Use the differential equation approach to find i(t) for t>0 in Fig. E7.4. 3 k t=0 2 k 6 mH i(t) Figure E7.4 2 k 12 V + – + 6V ANSWER: i(t)= -6 -2+6e-t/5⫻10 mA. 305 irwin07_296-368hr.qxd 306 28-07-2010 CHAPTER 7 11:34 Page 306 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S The Step-by-Step Approach In the previous analysis technique, we derived the differential equation for the capacitor voltage or inductor current, solved the differential equation, and used the solution to find the unknown variable in the network. In the very methodical technique that we will now describe, we will use the fact that Eq. (7.11) is the form of the solution and we will employ circuit analysis to determine the constants K1 , K2 , and . From Eq. (7.11) we note that as t S q, e-at S 0 and x(t) = K1 . Therefore, if the circuit is solved for the variable x(t) in steady state (i.e., t S q) with the capacitor replaced by an open circuit [v is constant and therefore i = C(dv兾dt) = 0] or the inductor replaced by a short circuit [i is constant and therefore v = L(di兾dt) = 0], then the variable x(t) = K1 . Note that since the capacitor or inductor has been removed, the circuit is a dc circuit with constant sources and resistors, and therefore only dc analysis is required in the steady-state solution. The constant K2 in Eq. (7.11) can also be obtained via the solution of a dc circuit in which a capacitor is replaced by a voltage source or an inductor is replaced by a current source. The value of the voltage source for the capacitor or the current source for the inductor is a known value at one instant of time. In general, we will use the initial condition value since it is generally the one known, but the value at any instant could be used. This value can be obtained in numerous ways and is often specified as input data in a statement of the problem. However, a more likely situation is one in which a switch is thrown in the circuit and the initial value of the capacitor voltage or inductor current is determined from the previous circuit (i.e., the circuit before the switch is thrown). It is normally assumed that the previous circuit has reached steady state, and therefore the voltage across the capacitor or the current through the inductor can be found in exactly the same manner as was used to find K1 . Finally, the value of the time constant can be found by determining the Thévenin equivalent resistance at the terminals of the storage element. Then = RTh C for an RC circuit, and = L兾RTh for an RL circuit. Let us now reiterate this procedure in a step-by-step fashion. Problem-Solving Strategy Using the Step-byStep Approach Step 1. We assume a solution for the variable x(t) of the form x(t) = K1 + K2 e-t兾. Step 2. Assuming that the original circuit has reached steady state before a switch was thrown (thereby producing a new circuit), draw this previous circuit with the capacitor replaced by an open circuit or the inductor replaced by a short circuit. Solve for the voltage across the capacitor, vC(0-), or the current through the inductor, iL(0-), prior to switch action. Step 3. Recall from Chapter 6 that voltage across a capacitor and the current flowing through an inductor cannot change in zero time. Draw the circuit valid for t = 0+ with the switches in their new positions. Replace a capacitor with a voltage source vC(0+) = vC(0-) or an inductor with a current source of value iL(0+) = iL(0-). Solve for the initial value of the variable x(0+). Step 4. Assuming that steady state has been reached after the switches are thrown, draw the equivalent circuit, valid for t 7 5, by replacing the capacitor by an open circuit or the inductor by a short circuit. Solve for the steady-state value of the variable x(t)|t 7 5 ⬟ x(q) Step 5. Since the time constant for all voltages and currents in the circuit will be the same, it can be obtained by reducing the entire circuit to a simple series circuit containing a voltage source, resistor, and a storage element (i.e., capacitor or inductor) by forming a simple Thévenin equivalent circuit at the terminals of irwin07_296-368hr.qxd 28-07-2010 11:34 Page 307 SECTION 7.2 F I R S T- O R D E R C I R C U I T S 307 the storage element. This Thévenin equivalent circuit is obtained by looking into the circuit from the terminals of the storage element. The time constant for a circuit containing a capacitor is = RTh C, and for a circuit containing an inductor it is = L兾RTh . Step 6. Using the results of steps 3, 4, and 5, we can evaluate the constants in step 1 as x(0+) = K1 + K2 x(q) = K1 Therefore, K1 = x(q), K2 = x(0+) - x(q), and hence the solution is x(t) = x(q) + [x(0+) - x(q)]e-t兾 Keep in mind that this solution form applies only to a first-order circuit having dc sources. If the sources are not dc, the forced response will be different. Generally, the forced response is of the same form as the forcing functions (sources) and their derivatives. Consider the circuit shown in Fig. 7.6a. The circuit is in steady state prior to time t=0, when the switch is closed. Let us calculate the current i(t) for t 7 0. Step 1. i(t) is of the form K1 + K2 e-t兾. Step 2. The initial voltage across the capacitor is calculated from Fig. 7.6b as vC (0-) = 36 - (2)(2) = 32 V Step 3. The new circuit, valid only for t = 0+, is shown in Fig. 7.6c. The value of the voltage source that replaces the capacitor is vC(0-) = vC(0+) = 32 V. Hence, 32 6k 16 mA = 3 i(0+) = Step 4. The equivalent circuit, valid for t 7 5, is shown in Fig. 7.6d. The current i(q) caused by the 36-V source is 36 2k + 6k 9 = mA 2 i(q) = Step 5. The Thévenin equivalent resistance, obtained by looking into the open-circuit terminals of the capacitor in Fig. 7.6e, is RTh = (2k)(6k) 2k + 6k = 3 k⍀ 2 Therefore, the circuit time constant is = RTh C 3 = a b A103 B(100)A10-6 B 2 = 0.15 s EXAMPLE 7.3 SOLUTION irwin07_296-368hr.qxd 308 28-07-2010 CHAPTER 7 11:34 Page 308 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 2 k 6 k i(t) 2 k 4 k 6 k i(0–) 4 k + 36 V ± – t=0 100 F 12 V ± – ± – vC(0–) 36 V - (b) (a) 2 k 36 V ± – 6 k ± – i(0±) 32 V 2 k 4 k 12 V ± – ± – 12 V ± – t=0– 6 k i(∞) 36 V 4 k 12 V ± – (d) t=∞ (c) t=0± i(t) (mA) 2 k 6 k 4 k 16 — 3 9 — 2 RTh 2 0 0.1 (e) 0.2 0.3 0.4 t(s) (f) Figure 7.6 Analysis of an RC transient circuit with a constant forcing function. Step 6. K1 = i(q) = 9 mA 2 K2 = i(0+) - i(q) = i(0+) - K1 = 9 16 3 2 = 5 mA 6 Therefore, i(t) = 36 5 + e-t兾0.15 mA 8 6 The plot is shown in Fig. 7.7 and can be compared to the sketch in Fig. 7.6f. Examination of Fig. 7.6f indicates once again that although the voltage across the capacitor is continuous at t=0, the current i(t) in the 6-k⍀ resistor jumps at t=0 from 2 mA to 5 1兾3 mA, and finally decays to 4 1兾2 mA. irwin07_296-368hr.qxd 28-07-2010 11:34 Page 309 SECTION 7.2 F I R S T- O R D E R C I R C U I T S 309 Figure 7.7 Plot for Example 7.3. The circuit shown in Fig. 7.8a is assumed to have been in a steady-state condition prior to switch closure at t=0. We wish to calculate the voltage v(t) for t 7 0. SOLUTION Step 1. v(t) is of the form K1 + K2 e-t兾. Step 2. In Fig. 7.8b we see that iL (0-) = 4 + = 6 24 a b (6)(3) 6 + 3 6 + 3 8 A 3 Step 3. The new circuit, valid only for t = 0+, is shown in Fig. 7.8c, which is equivalent to the circuit shown in Fig. 7.8d. The value of the current source that replaces the inductor is iL(0-) = iL(0+) = 8兾3 A. The node voltage v1(0+) can be determined from the circuit in Fig. 7.8d using a single-node equation, and v(0+) is equal to the difference between the source voltage and v1(0+). The equation for v1(0+) is v1(0+) - 24 4 v1(0+) + EXAMPLE 6 + v1(0+) 8 + = 0 3 12 or v1(0+) = 20 V 3 7.4 irwin07_296-368hr.qxd 310 28-07-2010 CHAPTER 7 11:34 Page 310 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 12 + v(t)- 12 1 4H 4 1 iL(0–) 4 24 V ± – 6 t=0 2 24 V ± – 6 2 (b) t=0– (a) 12 k + v(0±)4 24 V ± – + v(0±)- v1(0±) 1 4 8 —A 3 6 2 24 V ± – 12 6 8 —A 3 (c) t=0± 2 (d) t=0± 12 + 1 12 v(∞) - 1 4 1 4 24 V ± – 6 2 6 2 RTh (e) t=∞ (f) v(t) (V) 24 17.33 16 0 1 2 3 4 5 t (s) (g) Figure 7.8 Analysis of an RL transient circuit with a constant forcing function. irwin07_296-368hr.qxd 28-07-2010 11:34 Page 311 SECTION 7.2 F I R S T- O R D E R C I R C U I T S Then v(0+) = 24 - v1(0+) = 52 V 3 Step 4. The equivalent circuit for the steady-state condition after switch closure is given in Fig. 7.8e. Note that the 6-, 12-, 1-, and 2-⍀ resistors are shorted, and therefore v(q)=24 V. Step 5. The Thévenin equivalent resistance is found by looking into the circuit from the inductor terminals. This circuit is shown in Fig. 7.8f. Note carefully that RTh is equal to the 4-, 6-, and 12-⍀ resistors in parallel. Therefore, RTh=2 ⍀, and the circuit time constant is = 4 L = = 2s RTh 2 Step 6. From the previous analysis we find that K1 = v(q) = 24 K2 = v(0+) - v(q) = - 20 3 and hence that v(t) = 24 - 20 -t兾2 V e 3 From Fig. 7.8b we see that the value of v(t) before switch closure is 16 V. This value jumps to 17.33 V at t=0. A plot of this function for t>0 is shown in Fig. 7.9. Figure 7.9 Plot for Example 7.4. 311 irwin07_296-368hr.qxd 312 28-07-2010 CHAPTER 7 11:34 Page 312 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Learning Assessments E7.5 Consider the network in Fig. E7.5. The switch opens at t = 0. Find vo(t) for t 7 0. 1 + 2 ANSWER: vo(t) = 24 1 + e-(5兾8)t V. 5 5 t=0 12 V ± – 2F 2 vo(t) 2 ± – 8V - Figure E7.5 E7.6 Consider the network in Fig. E7.6. If the switch opens at t = 0, find the output voltage ANSWER: vo(t) for t 7 0. vo(t) = 6 2 10 -2t e V. 3 + 2H t=0 12 V ± – 2 vo(t) 2 + 4V - Figure E7.6 E7.7 Find v0(t) for t>0 in Fig. E7.7 using the step-by-step method. 4 k 24 V + – ANSWER: vo(t)=-3.33e-t/0.06 V. t=0 + vo(t) - 6 3 6 5 F 4 k 12 V Figure E7.7 + – irwin07_296-368hr.qxd 28-07-2010 11:34 Page 313 SECTION 7.2 F I R S T- O R D E R C I R C U I T S E7.8 Find i0(t) for t>0 Fig. E7.8 using the step-by-step method. 4 313 ANSWER: i0(t)=2.1-0.6e-t/0.001 A. 3 t=0 6 100 mH + – 12 V 3 3A io(t) Figure E7.8 The circuit shown in Fig. 7.10a has reached steady state with the switch in position 1. At time t=0 the switch moves from position 1 to position 2. We want to calculate vo(t) for t>0. Step 1. vo(t) is of the form K1 + K2 e-t兾. Step 2. Using the circuit in Fig. 7.10b, we can calculate iL(0-) iA = 12 = 3A 4 Then iL(0-) = 12 + 2iA 18 = = 3A 6 6 Step 3. The new circuit, valid only for t = 0+, is shown in Fig. 7.10c. The value of the current source that replaces the inductor is iL(0-) = iL(0+) = 3 A.Because of the current source vo(0+) = (3)(6) = 18 V Step 4. The equivalent circuit, for the steady-state condition after switch closure, is given in Fig. 7.10d. Using the voltages and currents defined in the figure, we can compute vo(q) in a variety of ways. For example, using node equations we can find vo(q) from vB vB + 2i Aœ vB - 36 + + = 0 2 4 6 vB i Aœ = 4 vo(q) = vB + 2iAœ or, using loop equations, 36 = 2Ai1 + i2 B + 4i1 36 = 2Ai1 + i2 B + 6i2 - 2i1 vo(q) = 6i2 Using either approach, we find that vo(q)=27 V. EXAMPLE 7.5 SOLUTION irwin07_296-368hr.qxd 314 28-07-2010 CHAPTER 7 11:34 Page 314 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Step 5. The Thévenin equivalent resistance can be obtained via voc and isc because of the presence of the dependent source. From Fig. 7.10e we note that iAfl = 36 = 6A 2 + 4 Therefore, voc = (4)(6) + 2(6) = 36 V 2 t=0 iL(0–) 3H 2 + 1 36 V ± – 12 V 4 6 vo(t) - ± – + iA + 4 12 V 2iA + iA 2iA (b) t=0– 2 3A vB + 4 36 V 6 + + vo(0±) - ± – iA 4 36 V ± – i1 36 V isc 6 ± – 4 36 V i"A 2i'A (d) t= 2 + ± – i A'" i A'" (e) Figure 7.10 Analysis of an RL transient circuit containing a dependent source. 6 isc + 2i"A (f) vo(∞) - i2 i'A + voc +– 4 6 2iA (c) t=0± 2 vo(0–) - ± – (a) 2 6 '" 2i A irwin07_296-368hr.qxd 28-07-2010 11:34 Page 315 SECTION 7.2 F I R S T- O R D E R C I R C U I T S 315 From Fig. 7.10f we can write the following loop equations: Ô 36 = 2AiÔ A + isc B + 4i A Ô 36 = 2AiÔ A + isc B + 6isc - 2i A Solving these equations for isc yields isc = 9 A 2 Therefore, RTh = voc 36 = 8⍀ = isc 9兾2 Hence, the circuit time constant is = 3 L = s RTh 8 Step 6. Using the information just computed, we can derive the final equation for vo(t): K1 = vo(q) = 27 K2 = vo(0+) - vo(q) = 18 - 27 = -9 Therefore, vo(t) = 27 - 9e-t兾(3兾8) V Learning Assessments E7.9 If the switch in the network in Fig. E7.9 closes at t = 0, find vo(t) for t 7 0. ANSWER: vo(t) = 24 + 36e-(t兾12) V. 24 V 4 + + 2 vA 2F + 3A t=0 4 vA + vo(t) - - Figure E7.9 E7.10 Find io(t) for t>0 in Fig. E7.10 using the step-by-step method. ANSWER: io(t) = 1.5 + 0.2143e-(t兾0.7) mA. t=0 iA 6 k 6 k 3 k 6 k io(t) 12 V + - Figure E7.10 100 F 2iA 4 k 2F irwin07_296-368hr.qxd 316 28-07-2010 CHAPTER 7 11:34 Page 316 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S At this point, it is appropriate to state that not all switch action will always occur at time t=0. It may occur at any time t0 . In this case the results of the step-by-step analysis yield the following equations: x At0 B = K1 + K2 x (q) = K1 and x(t) = x(q) + C x At0 B - x(q)D e-At - t0B兾 t 7 t0 The function is essentially time-shifted by t0 seconds. Finally, note that if more than one independent source is present in the network, we can simply employ superposition to obtain the total response. PULSE RESPONSE Thus far we have examined networks in which a voltage or current source is suddenly applied. As a result of this sudden application of a source, voltages or currents in the circuit are forced to change abruptly. A forcing function whose value changes in a discontinuous manner or has a discontinuous derivative is called a singular function. Two such singular functions that are very important in circuit analysis are the unit impulse function and the unit step function. We will defer a discussion of the unit impulse function until a later chapter and concentrate on the unit step function. The unit step function is defined by the following mathematical relationship: u(t) = b 0 1 t 6 0 t 7 0 In other words, this function, which is dimensionless, is equal to zero for negative values of the argument and equal to 1 for positive values of the argument. It is undefined for a zero argument where the function is discontinuous. A graph of the unit step is shown in Fig. 7.11a. 2 Figure 7.11 Graphs and models of the unit step function. 1 1 Vo ± – t=0 t 0 (a) (b) t=0 Vou(t) ± – Io (c) (d) u(t-t0) 1 Iou(t-t0) 0 (e) t0 (f) t irwin07_296-368hr.qxd 28-07-2010 11:34 Page 317 SECTION 7.2 F I R S T- O R D E R C I R C U I T S The unit step is dimensionless, and therefore a voltage step of Vo volts or a current step of Io amperes is written as Vo u(t) and Iou(t), respectively. Equivalent circuits for a voltage step are shown in Figs. 7.11b and c. Equivalent circuits for a current step are shown in Figs. 7.11d and e. If we use the definition of the unit step, it is easy to generalize this function by replacing the argument t by t-t0 . In this case 0 u At - t0 B = b 1 v(t) t 6 t0 t 7 t0 A A graph of this function is shown in Fig. 7.11f. Note that u At - t0 B is equivalent to delaying u(t) by t0 seconds, so that the abrupt change occurs at time t=t0 . Step functions can be used to construct one or more pulses. For example, the voltage pulse shown in Fig. 7.12a can be formulated by initiating a unit step at t=0 and subtracting one that starts at t=T, as shown in Fig. 7.12b. The equation for the pulse is (a) v(t) Au(t) A T If the pulse is to start at t=t0 and have width T, the equation would be t v(t) = AEu At - t0 B - uC t - At0 + TB D F Using this approach, we can write the equation for a pulse starting at any time and ending at any time. Similarly, using this approach, we could write the equation for a series of pulses, called a pulse train, by simply forming a summation of pulses constructed in the manner illustrated previously. The following example will serve to illustrate many of the concepts we have just presented. –A –Au(t-T) (b) Figure 7.12 Construction of a pulse via two step functions. Consider the circuit shown in Fig. 7.13a. The input function is the voltage pulse shown in Fig. 7.13b. Since the source is zero for all negative time, the initial conditions for the v(t) ± – EXAMPLE 7.6 4 k + + vC(t) - 100 F v(t) (V) 9 vo(t) 8 k 0 0.3 (a) t(s) (b) 4 k 6 k 4 k + + vc(0)=0 8 k vo(0±)=0 ± – 9V 8 k t=0 6 k vo(∞) - - (c) t T v(t) = ACu(t) - u(t - T)D 6 k 317 (d) 4 k vo(t) (V) 4 8 k RTh 2.11 0 (e) 0.3 t(s) (f) Figure 7.13 Pulse response of a network. irwin07_296-368hr.qxd 318 28-07-2010 CHAPTER 7 11:34 Page 318 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S network are zero C i.e., vC(0-) = 0D. The response vo(t) for 0 6 t 6 0.3 s is due to the application of the constant source at t=0 and is not influenced by any source changes that will occur later. At t = 0.3 s the forcing function becomes zero, and therefore vo(t) for t 7 0.3 s is the source-free or natural response of the network. Let us determine the expression for the voltage vo(t). SOLUTION Since the output voltage vo(t) is a voltage division of the capacitor voltage, and the initial voltage across the capacitor is zero, we know that vo(0+) = 0, as shown in Fig. 7.13c. If no changes were made in the source after t=0, the steady-state value of vo(t) Ci.e., vo(q)D due to the application of the unit step at t=0 would be 9 (8k) 6k + 4k + 8k = 4V vo(q) = as shown in Fig. 7.13d. The Thévenin equivalent resistance is RTh = (6k)(12k) 6k + 12k = 4 k⍀ as illustrated in Fig. 7.13e. Therefore, the circuit time constant is = RTh C = (4)A103 B(100)A10-6 B = 0.4 s Therefore, the response vo(t) for the period 0<t<0.3 s is vo(t) = 4 - 4e-t兾0.4 V 0<t<0.3 s The capacitor voltage can be calculated by realizing that using voltage division, vo(t) = 2兾3 vC(t). Therefore, vC(t) = 3 A4 - 4e-t兾0.4 B V 2 Since the capacitor voltage is continuous, vC(0.3-) = vC(0.3+) and therefore, vo(0.3+) = 2 v (0.3-) 3 C = 4A1 - e-0.3兾0.4 B = 2.11 V Since the source is zero for t 7 0.3 s, the final value for vo(t) as t S q is zero. Therefore, the expression for vo(t) for t 7 0.3 s is vo(t) = 2.11e-(t - 0.3)兾0.4 V t>0.3 s The term e-(t - 0.3)兾0.4 indicates that the exponential decay starts at t=0.3 s. The complete solution can be written by means of superposition as vo(t) = 4A1 - e-t兾0.4 Bu(t) - 4A1 - e-(t - 0.3)兾0.4 Bu(t - 0.3) V irwin07_296-368hr.qxd 28-07-2010 11:34 Page 319 SECTION 7.3 SECOND-ORDER CIRCUITS 319 or, equivalently, the complete solution is 0 vo(t) = • 4A1 - e-t兾0.4 B V 2.11e-(t - 0.3)兾0.4 V t 6 0 0 6 t 6 0.3 s ¶ 0.3 s 6 t which in mathematical form is vo(t) = 4A1 - e-t兾0.4 B Cu(t) - u(t - 0.3) D + 2.11e-(t - 0.3)兾0.4 u(t - 0.3) V Note that the term Cu(t)-u(t-0.3)D acts like a gating function that captures only the part of the step response that exists in the time interval 0<t<0.3 s. The output as a function of time is shown in Fig. 7.13f. Learning Assessments ANSWER: vo(t) = 0 for t 6 0, 4A1 - e-(3兾2)t B V for 0 ⬉ t ⬉ 1 s, and 3.11e-(3兾2)(t - 1) V for 1 s<t. E7.11 The voltage source in the network in Fig. E7.11a is shown in Fig. E7.11b. The initial current in the inductor must be zero. (Why?) Determine the output voltage vo(t) for t 7 0. 2H 2 v(t) ± – + 2 2 v(t) (V) vo(t) 12 0 (a) 1 t(s) (b) Figure E7.11 7.3 T H E B A S I C C I R C U I T EQ U AT I O N To begin our development, let us consider the two basic RLC circuits shown in Fig. 7.14. We assume that energy may be initially stored in both the inductor and capacitor. The node equation for the parallel RLC circuit is Second-Order Circuits t 1 dv v + v(x) dx + iL At0 B + C = iS(t) R L 3t0 dt Similarly, the loop equation for the series RLC circuit is Ri + t 1 di i(x) dx + vC At0 B + L = vS(t) C 3t0 dt v(t) i(t) vC(t0) R + iL(t0) iS(t) R L C (a) Figure 7.14 Parallel and series RLC circuits. - C vS(t) ± – L (b) irwin07_296-368hr.qxd 320 28-07-2010 CHAPTER 7 11:34 Page 320 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Note that the equation for the node voltage in the parallel circuit is of the same form as that for the loop current in the series circuit. Therefore, the solution of these two circuits is dependent on solving one equation. If the two preceding equations are differentiated with respect to time, we obtain C d 2v 2 dt di S 1 dv v + = R dt L dt + and L d2i dt 2 + R dvS di i + = dt C dt Since both circuits lead to a second-order differential equation with constant coefficients, we will concentrate our analysis on this type of equation. T H E R E S P O N S E EQ U AT I O N S In concert with our development of the solution of a first-order differential equation that results from the analysis of either an RL or an RC circuit as outlined earlier, we will now employ the same approach here to obtain the solution of a second-order differential equation that results from the analysis of RLC circuits. As a general rule, for this case we are confronted with an equation of the form d2x(t) dt 2 + a1 dx(t) dt + a2 x(t) = f(t) 7.12 Once again we use the fact that if x(t)=xp(t) is a solution to Eq. (7.12), and if x(t)=xc(t) is a solution to the homogeneous equation d 2x(t) dt 2 + a1 dx(t) dt + a2 x(t) = 0 then x(t) = xp(t) + xc(t) is a solution to the original Eq. (7.12). If we again confine ourselves to a constant forcing function Ci.e., f(t)=AD, the development at the beginning of this chapter shows that the solution of Eq. (7.12) will be of the form x(t) = A + xc(t) a2 7.13 Let us now turn our attention to the solution of the homogeneous equation d2x(t) dt 2 + a1 dx(t) dt + a2 x(t) = 0 where a1 and a2 are constants. For simplicity we will rewrite the equation in the form d2x(t) dt 2 + 20 dx(t) dt + 20 x(t) = 0 7.14 where we have made the following simple substitutions for the constants a1 = 20 and a2 = 20 . Following the development of a solution for the first-order homogeneous differential equation earlier in this chapter, the solution of Eq. (7.14) must be a function whose first- and second-order derivatives have the same form, so that the left-hand side of Eq. (7.14) will become identically zero for all t. Again we assume that x(t) = Kest Substituting this expression into Eq. (7.14) yields s2Kest + 20 sKest + 20 Kest = 0 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 321 SECTION 7.3 SECOND-ORDER CIRCUITS Dividing both sides of the equation by Kest yields s2 + 20 s + 20 = 0 7.15 This equation is commonly called the characteristic equation; is called the exponential damping ratio, and 0 is referred to as the undamped natural frequency. The importance of this terminology will become clear as we proceed with the development. If this equation is satisfied, our assumed solution x(t) = Kest is correct. Employing the quadratic formula, we find that Eq. (7.15) is satisfied if -20 ; 24220 - 420 2 s = = -0 ; 0 22 - 1 7.16 Therefore, two values of s, s1 and s2, satisfy Eq. (7.15): s1 = -0 + 0 22 - 1 7.17 s2 = -0 - 0 22 - 1 In general, then, the complementary solution of Eq. (7.14) is of the form xc(t) = K1 es1 t + K2 es2 t 7.18 K1 and K2 are constants that can be evaluated via the initial conditions x(0) and dx(0)兾dt. For example, since x(t) = K1 es1 t + K2 es2 t then x(0) = K1 + K2 and dx(t) dt dx(0) 2 = t=0 dt = s1 K1 + s2 K2 Hence, x(0) and dx (0)兾dt produce two simultaneous equations, which when solved yield the constants K1 and K2 . Close examination of Eqs. (7.17) and (7.18) indicates that the form of the solution of the homogeneous equation is dependent on the value . For example, if >1, the roots of the characteristic equation, s1 and s2 , also called the natural frequencies because they determine the natural (unforced) response of the network, are real and unequal; if <1, the roots are complex numbers; and finally, if =1, the roots are real and equal. Let us now consider the three distinct forms of the unforced response—that is, the response due to an initial capacitor voltage or initial inductor current. Case 1, >1 This case is commonly called overdamped. The natural frequencies s1 and s2 are real and unequal; therefore, the natural response of the network described by the second-order differential equation is of the form xc(t) = K1 e- A0 - 0 2 2 - 1Bt + K2 e- A0 + 0 2 2 - 1Bt 7.19 where K1 and K2 are found from the initial conditions. This indicates that the natural response is the sum of two decaying exponentials. 321 irwin07_296-368hr.qxd 322 28-07-2010 CHAPTER 7 11:34 Page 322 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Case 2, <1 This case is called underdamped. Since <1, the roots of the characteristic equation given in Eq. (7.17) can be written as s1 = -0 + j0 21 - 2 = - + jd s2 = -0 - j0 21 - 2 = - - jd where j = 1-1 , = 0, and d = 0 21 - 2 . Thus, the natural frequencies are complex numbers (briefly discussed in the Appendix). The natural response is then of the form xc(t) = e-0 t AA 1 cos 0 21 - 2 t + A 2 sin 0 21 - 2 tB 7.20 where A1 and A2 , like K1 and K2 , are constants, which are evaluated using the initial conditions x(0) and dx(0)兾dt. This illustrates that the natural response is an exponentially damped oscillatory response. Case 3, =1 This case, called critically damped, results in s1 = s2 = -0 In the case where the characteristic equation has repeated roots, the general solution is of the form xc(t) = B1 e-0 t + B2 te-0 t 7.21 where B1 and B2 are constants derived from the initial conditions. It is informative to sketch the natural response for the three cases we have discussed: overdamped, Eq. (7.19); underdamped, Eq. (7.20); and critically damped, Eq. (7.21). Figure 7.15 graphically illustrates the three cases for the situations in which xc(0)=0. Note that the critically damped response peaks and decays faster than the overdamped response. The underdamped response is an exponentially damped sinusoid whose rate of decay is dependent on the factor . Actually, the terms ; e-0 t define what is called the envelope of the response, and the damped oscillations (i.e., the oscillations of decreasing amplitude) exhibited by the waveform in Fig. 7.15b are called ringing. Figure 7.15 Comparison of overdamped, critically damped, and underdamped responses. xc(t) xc(t) Critically damped Underdamped e–t Overdamped t t (a) (b) Learning Assessments E7.12 A parallel RLC circuit has the following circuit parameters: R=1 ⍀, L=2 H, and C=2 F. Compute the damping ratio and the undamped natural frequency of this network. E7.13 A series RLC circuit consists of R=2 ⍀, L=1 H, and a capacitor. Determine the type of response exhibited by the network if (a) C = 1兾2 F, (b) C=1 F, and (c) C=2 F. THE NETWORK RESPONSE ANSWER: = 0.5; 0 = 0.5 rad兾s. ANSWER: (a) underdamped; (b) critically damped; (c) overdamped. We will now analyze a number of simple RLC networks that contain both nonzero initial conditions and constant forcing functions. Circuits that exhibit overdamped, underdamped, and critically damped responses will be considered. irwin07_296-368hr.qxd 28-07-2010 11:34 Page 323 SECTION 7.3 SECOND-ORDER CIRCUITS 323 Problem-Solving Strategy Step 1. Write the differential equation that describes the circuit. Step 2. Derive the characteristic equation, which can be written in the form s2 + 20 s + 20 = 0, where is the damping ratio and 0 is the undamped natural frequency. Step 3. The two roots of the characteristic equation will determine the type of response. If the roots are real and unequal (i.e., >1), the network response is overdamped. If the roots are real and equal (i.e., =1), the network response is critically damped. If the roots are complex (i.e., <1), the network response is underdamped. Step 4. The damping condition and corresponding response for the aforementioned three cases outlined are as follows: 2 2 Overdamped: x(t) = K1 e-A0 - 0 2 - 1Bt + K2 e-A0 + 0 2 - 1Bt Critically damped: x(t) = B1 e-0 t + B2 te-0 t Underdamped: x(t) = e-t AA 1 cos d t + A 2 sin d tB, where = 0 , and d = 0 21 - 2 Step 5. Two initial conditions, either given or derived, are required to obtain the two unknown coefficients in the response equation. Second-Order Transient Circuits The following examples will demonstrate the analysis techniques. Consider the parallel RLC circuit shown in Fig. 7.16. The second-order differential equation that describes the voltage v(t) is d2v EXAMPLE 7.7 v 1 dv + + = 0 RC dt LC dt 2 v(t) Figure 7.16 + R L C iL(0) Parallel RLC circuit. vC(0) - A comparison of this equation with Eqs. (7.14) and (7.15) indicates that for the parallel RLC circuit the damping term is 1/2 RC and the undamped natural frequency is 1兾1LC . If the circuit parameters are R=2 ⍀, C = 1兾5 F, and L=5 H, the equation becomes d2v dt 2 + 2.5 dv + v = 0 dt Let us assume that the initial conditions on the storage elements are iL(0)=-1 A and vC(0)=4 V. Let us find the node voltage v(t) and the inductor current. The characteristic equation for the network is s2 + 2.5s + 1 = 0 and the roots are s1 = -2 s2 = -0.5 SOLUTION irwin07_296-368hr.qxd 324 28-07-2010 CHAPTER 7 11:34 Page 324 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Since the roots are real and unequal, the circuit is overdamped, and v(t) is of the form v(t) = K1 e-2t + K2 e-0.5t The initial conditions are now employed to determine the constants K1 and K2 . Since v(t)=vC(t), vC(0) = v(0) = 4 = K1 + K2 The second equation needed to determine K1 and K2 is normally obtained from the expression dv(t) dt = -2K1 e-2t - 0.5K2 e-0.5t However, the second initial condition is not dv(0)兾dt. If this were the case, we would simply evaluate the equation at t=0. This would produce a second equation in the unknowns K1 and K2 . We can, however, circumvent this problem by noting that the node equation for the circuit can be written as v(t) dv(t) C + dt R + iL(t) = 0 or iL(t) -1 = v(t) RC C dv(t) dt At t=0, dv(0) dt = -1 1 v(0) - iL(0) RC C = -2.5(4) - 5(-1) = -5 However, since dv(t) dt = -2K1 e-2t - 0.5K2 e-0.5t then when t=0 -5 = -2K1 - 0.5K2 This equation, together with the equation 4 = K1 + K2 produces the constants K1 = 2 and K2 = 2 . Therefore, the final equation for the voltage is v(t) = 2e-2t + 2e-0.5t V Note that the voltage equation satisfies the initial condition v(0)=4 V. The response curve for this voltage v(t) is shown in Fig. 7.17. Figure 7.17 Overdamped response. v(t) (V) 4.8 4.2 3.6 3.0 2.4 1.8 1.2 0.6 0.0 t(s) 0.0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 325 SECTION 7.3 SECOND-ORDER CIRCUITS 325 The inductor current is related to v(t) by the equation iL(t) = 1 v(t) dt L 3 Substituting our expression for v(t) yields iL(t) = or 1 C 2e-2t + 2e-0.5t D dt 5 3 iL(t) = - 1 -2t 4 e - e-0.5t A 5 5 Note that in comparison with the RL and RC circuits, the response of this RLC circuit is controlled by two time constants. The first term has a time constant of 1兾2 s, and the second term has a time constant of 2 s. The series RLC circuit shown in Fig. 7.18 has the following parameters: C=0.04 F, L=1 H, R=6 ⍀, iL(0)=4 A, and vC(0)=-4 V. The equation for the current in the circuit is given by the expression 2 d i dt 2 EXAMPLE 7.8 R di i + = 0 L dt LC + A comparison of this equation with Eqs. (7.14) and (7.15) illustrates that for a series RLC circuit the damping term is R/2L and the undamped natural frequency is 1兾1LC . Substituting the circuit element values into the preceding equation yields d2i dt 2 + 6 di + 25i = 0 dt Let us determine the expression for both the current and the capacitor voltage. SOLUTION The characteristic equation is then s2 + 6s + 25 = 0 and the roots are s1 = -3 + j4 s2 = -3 - j4 Since the roots are complex, the circuit is underdamped, and the expression for i(t) is i(t) = K1 e-3t cos 4t + K2 e-3t sin 4t Using the initial conditions, we find that i(0)=4=K1 i(t) iL(0) Figure 7.18 Series RLC circuit. L R + C vC(0) - irwin07_296-368hr.qxd 326 28-07-2010 CHAPTER 7 11:34 Page 326 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S and di = -4K1 e-3t sin 4t - 3K1 e-3t cos 4t dt + 4K2 e-3t cos 4t - 3K2 e-3t sin 4t and thus di(0) dt = -3K1 + 4K2 Although we do not know di(0)兾dt, we can find it via KVL. From the circuit we note that Ri(0) + L di(0) dt + vC(0) = 0 or di(0) dt vC(0) R i(0) L L 4 6 = - (4) + 1 1 = -20 = - Therefore, -3K1 + 4K2 = -20 and since K1=4, K2=–2, the expression then for i(t) is i(t) = 4e-3t cos 4t - 2e-3t sin 4t A Note that this expression satisfies the initial condition i(0)=4. The voltage across the capacitor could be determined via KVL using this current: Ri(t) + L di(t) dt + vC(t) = 0 or vC(t) = -Ri(t) - L di(t) dt Substituting the preceding expression for i(t) into this equation yields vC(t) = -4e-3t cos 4t + 22e-3t sin 4t V Note that this expression satisfies the initial condition vC(0)=–4 V. A plot of the function is shown in Fig. 7.1.9: Figure 7.19 Underdamped response. v(t) (V) 10.0 8.0 6.0 4.0 2.0 0 –2.0 –4.0 t(s) 0.0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 327 SECTION 7.3 SECOND-ORDER CIRCUITS Let us examine the circuit in Fig. 7.20, which is slightly more complicated than the two we have already considered. The two equations that describe the network are L di(t) dt EXAMPLE SOLUTION + R1 i(t) + v(t) = 0 v(t) dv(t) i(t) = C dt + R2 Substituting the second equation into the first yields d2v dt 2 + a R1 dv R1 + R2 1 + b + v = 0 R2 C L dt R2 LC If the circuit parameters and initial conditions are 1 F 8 R1 = 10 ⍀ C = R2 = 8 ⍀ L = 2H vC(0) = 1 V iL(0) = 1 A 2 the differential equation becomes d2 v dt 2 + 6 dv + 9v = 0 dt We wish to find expressions for the current i(t) and the voltage v(t). The characteristic equation is then s2 + 6s + 9 = 0 and hence the roots are s1 = -3 s2 = -3 Since the roots are real and equal, the circuit is critically damped. The term v(t) is then given by the expression v(t) = K1 e-3t + K2 te-3t Since v(t)=vC(t), v(0) = vC(0) = 1 = K1 i(t) Figure 7.20 Series-parallel RLC circuit. R1 iL(0) L + + vC(0) - C v(t) - R2 327 7.9 irwin07_296-368hr.qxd 328 28-07-2010 CHAPTER 7 11:34 Page 328 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S In addition, dv(t) dt = -3K1 e-3t + K2 e-3t - 3K2 te-3t However, i(t) dv(t) dt = C v(t) - R2 C Setting these two expressions equal to one another and evaluating the resultant equation at t=0 yields 1 1兾2 - = -3K1 + K2 1 1兾8 3 = -3K1 + K2 K1=1, K2=6, and the expression for v(t) is v(t) = e-3t + 6te-3t V Note that the expression satisfies the initial condition v(0)=1. The current i(t) can be determined from the nodal analysis equation at v(t): i(t) = C v(t) dv(t) dt + R2 Substituting v(t) from the preceding equation, we find i(t) = 1 1 C -3e-3t + 6e-3t - 18te-3t D + C e-3t + 6te-3t D 8 8 or i(t) = 3 1 -3t e - te-3t A 2 2 If this expression for the current is employed in the circuit equation, v(t) = -L di(t) dt - R1 i(t) we obtain v(t) = e-3t + 6te-3t V which is identical to the expression derived earlier. A plot of this critically damped function is shown in Fig. 7.21. Figure 7.21 Critically damped response. v(t) (V) 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 –0.2 t(s) 0.0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7 3.0 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 329 SECTION 7.3 SECOND-ORDER CIRCUITS 329 Learning Assessments E7.14 The switch in the network in Fig. E7.14 opens at t=0. Find i(t) for t>0. 6 3 t=0 1F ± – i(t) Figure E7.14 ANSWER: i(t) = -2e - t兾2 + 4e-t A. 2H 12 V E7.15 The switch in the network in Fig. E7.15 moves from position 1 to position 2 at t=0. Find vo(t) for t>0. 1 ANSWER: vo(t) = 2Ae-t - 3e-3t B V. 2 t=0 2 —F 3 1 —H 2 2A + vo(t) 2 - Figure E7.15 E7.16 Find vC(t) for t>0 in Fig. E7.16. ANSWER: vC(t)=-2e–2t - 1 vc(t) + cos t-1.5e–2t sin t+24 V. 0.8 F t=0 24 V 0.25 H 5 12 V + – +V – S2 Figure E7.16 Consider the circuit shown in Fig. 7.22. This circuit is the same as the one analyzed in Example 7.8, except that a constant forcing function is present. The circuit parameters are the same as those used in Example 7.8: C = 0.04 F iL(0) = 4 A L = 1H vC(0) = -4 V R = 6⍀ EXAMPLE 7.10 irwin07_296-368hr.qxd 330 28-07-2010 CHAPTER 7 11:34 Page 330 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Figure 7.22 iL(0) R Series RLC circuit with a step function input. L 12u(t) V ± – + C vC (0) - We want to find an expression for vC(t) for t>0. SOLUTION From our earlier mathematical development we know that the general solution of this problem will consist of a particular solution plus a complementary solution. From Example 7.8 we know that the complementary solution is of the form K3 e-3t cos 4t + K4 e-3t sin 4t. The particular solution is a constant, since the input is a constant and therefore the general solution is vC(t) = K3 e-3t cos 4t + K4 e-3t sin 4t + K5 An examination of the circuit shows that in the steady state, the final value of vC(t) is 12 V, since in the steady-state condition, the inductor is a short circuit and the capacitor is an open circuit. Thus, K5 = 12. The steady-state value could also be immediately calculated from the differential equation. The form of the general solution is then vC(t) = K3 e-3t cos 4t + K4 e-3t sin 4t + 12 The initial conditions can now be used to evaluate the constants K3 and K4 : vC(0) = -4 = K3 + 12 -16 = K3 Since the derivative of a constant is zero, the results of Example 7.8 show that dvC(0) i(0) = dt C = 100 = -3K3 + 4K4 and since K3 = -16, K4 = 13. Therefore, the general solution for vC(t) is vC(t) = 12 - 16e-3t cos 4t + 13e-3t sin 4t V Note that this equation satisfies the initial condition vC(0)=–4 and the final condition vC(q)=12 V. EXAMPLE 7.11 Let us examine the circuit shown in Fig. 7.23. A close examination of this circuit will indicate that it is identical to that shown in Example 7.9 except that a constant forcing function is present. We assume the circuit is in steady state at t = 0-. The equations that describe the circuit for t 7 0 are L di(t) dt + R1 i(t) + v(t) = 24 i(t) = C v(t) dv(t) dt + R2 Combining these equations, we obtain d2v(t) dt 2 + a R1 dv(t) R1 + R2 1 24 + + v(t) = b R2 C L dt R2 LC LC If the circuit parameters are R1 = 10 ⍀, R2 = 2 ⍀, L = 2 H, and C = 1兾4 F, the differential equation for the output voltage reduces to d2v(t) dt 2 + 7 dv(t) dt + 12v(t) = 48 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 331 SECTION 7.3 SECOND-ORDER CIRCUITS 331 Figure 7.23 t=0 24 V + iL(0) ± – ± – Series-parallel RLC circuit with a constant forcing function. R1 i(t) L + vC(0) - 12 V R2 C v(t) - Let us determine the output voltage v(t). The characteristic equation is s2 + 7s + 12 = 0 and hence the roots are s1 = -3 s2 = -4 The circuit response is overdamped, and therefore the general solution is of the form v(t) = K1 e-3t + K2 e-4t + K3 The steady-state value of the voltage, K3 , can be computed from Fig. 7.24a. Note that v(q) = 4 V = K3 The initial conditions can be calculated from Figs. 7.24b and c, which are valid at t = 0- and t = 0+, respectively. Note that v(0+) = 2 V and, hence, from the response equation v(0+) = 2 V = K1 + K2 + 4 -2 = K1 + K2 Fig. 7.24c illustrates that i(0+) = 1. From the response equation we see that dv(0) dt i(∞)=2 A = -3K1 - 4K2 10 i(0–)=1 A 10 + 24 V ± – 2 + v(∞)=4 V 12 V ± – + vC(0–)=2 V - 2 - - (a) t=∞ i(0±)=1 A (b) t=0– 10 i(0±) + iC(0±) 24 V ± – 2V ± – 2 v(0±)=2 V - (c) t=0+ Figure 7.24 v(0–)=2 V Equivalent circuits at t = q , t = 0- , and t = 0+ for the circuit in Fig. 7.23. irwin07_296-368hr.qxd 332 28-07-2010 CHAPTER 7 11:34 Page 332 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S and since i(0) dv(0) dt = C v(0) - R2 C = 4 - 4 = 0 then 0 = -3K1 - 4K2 Solving the two equations for K1 and K2 yields K1 = -8 and K2 = 6. Therefore, the general solution for the voltage response is v(t) = 4 - 8e-3t + 6e-4t V Note that this equation satisfies both the initial and final values of v(t). Learning Assessments E7.17 The switch in the network in Fig. E7.17 moves from position 1 to position 2 at t=0. Compute io(t) for t>0 and use this current to determine vo(t) for t>0. io(t) 1 6 + 2H 2 t=0 18 1 —F 36 24 V vo(t) ± – ± – ± – 4V 12 V - Figure E7.17 E7.18 Find i(t) for t>0 in Fig. E7.18. ANSWER: i(t)=0.4144e–17.07t – 2.414e–2.93t+3 A. 10 t=0 i(t) 10 10 V + – 2H 20 V Figure E7.18 ANSWER: 11 14 -6t e A; io(t) = - e-3t + 6 6 vo(t)=12+18io(t) V. + – 0.01F irwin07_296-368hr.qxd 28-07-2010 11:34 Page 333 SECTION 7.4 A P P L I C AT I O N E X A M P L E S 333 7.4 There are a wide variety of applications for transient circuits. The following examples demonstrate some of them. Application Examples Let us return to the camera flash circuit, redrawn in Fig. 7.25, which was discussed in the introduction of this chapter. The Xenon flash has the following specifications: Voltage required for successful flash: b APPLICATION EXAMPLE 7.12 minimum 50 V maximum 70 V Equivalent resistance: 80 ⍀ For this particular application, a time constant of 1 ms is required during flash time. In addition, to minimize the physical size of the circuit, the resistor R1 must dissipate no more than 100-mW peak power. We want to determine values for VS , CF , and R1 . Furthermore, we wish to determine the recharge time, the flash bulb’s voltage, current, power, and total energy dissipated during the flash. R1 VS ± – iB(t) A model for a camera flash charging circuit. + + vCF(t) Figure 7.25 CF - vB(t) RB - We begin by selecting the source voltage, VS . Since the capacitor is applied directly to the Xenon bulb during flash, and since at least 50 V is required to flash, we should set VS higher than 50 V. We will somewhat arbitrarily split the difference in the bulb’s required voltage range and select 60 V for VS . Now we consider the time constant during the flash time. From Fig. 7.25, during flash, the time constant is simply F = RB CF 7.22 Given tF = 1 ms and RB = 80 ⍀, we find CF = 12.5 F. Next, we turn to the value of R1 . At the beginning of the charge time, the capacitor voltage is zero and both the current and power in R1 are at their maximum values. Setting the power to the maximum allowed value of 100 mW, we can write PRmax = V2S 3600 = = 0.1 R1 R1 7.23 and find that R1 = 36 k⍀. The recharge time is the time required for the capacitor to charge from zero up to at least 50 V. At that point the flash can be successfully discharged. We will define t = 0 as the point at which the switch moves from the bulb back to R1 . At t = 0, the capacitor voltage is zero, and at t = q, the capacitor voltage is 60 V; the time constant is simply R1 CF . The resulting equation for the capacitor voltage during recharge is vCF(t) = K1 + K2 e-t兾 = 60 - 60e-t兾R1 CF V 7.24 At t = tcharge , vCF(t) = 50 V. Substituting this and the values of R1 and CF into Eq. (7.24) yields a charge time of tcharge = 806 ms—just less than a second. As a point of interest, let us SOLUTION • irwin07_296-368hr.qxd 334 28-07-2010 CHAPTER 7 11:34 Page 334 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S reconsider our choice for VS . What happens if VS is decreased to only 51 V? First, from Eq. (7.23), R1 changes to 26.01 k⍀. Second, from Eq. (7.24), the charge time increases only slightly to 1.28 s. Therefore, it appears that selection of VS will not have much effect on the flash unit’s performance, and thus there exists some flexibility in the design. Finally, we consider the waveforms for the flash bulb itself. The bulb and capacitor voltage are the same during flash and are given by the decaying exponential function vB(t) = 60e-1000t V 7.25 where the time constant is defined in Eq. (7.22), and we have assumed that the capacitor is allowed to charge fully to VS (i.e., 60 V). Since the bulb’s equivalent resistance is 80 ⍀, the bulb current must be iB(t) = 60e-1000t = 750e-1000t mA 80 7.26 As always, the power is the v-i product. pB(t) = vB(t)iB(t) = 45e-2000t W 7.27 Finally, the total energy consumed by the bulb during flash is only q wB(t) = APPLICATION EXAMPLE 7.13 30 q pB(t) dt = 30 45e-2000t dt = 45 -2000t 0 45 2 = e = 22.5 mJ 2000 2000 q One very popular application for inductors is storing energy in the present for release in the future. This energy is in the form of a magnetic field, and current is required to maintain the field. In an analogous situation, the capacitor stores energy in an electric field, and a voltage across the capacitor is required to maintain it. As an application of the inductor’s energy storage capability, let us consider the high-voltage pulse generator circuit shown in Fig. 7.26. This circuit is capable of producing high-voltage pulses from a small dc voltage. Let’s see if this circuit can produce an output voltage peak of 500 V every 2 ms, that is, 500 times per second. pos. 1 Figure 7.26 A simple high-voltage pulse generator. Vin ± 5V – pos. 2 i(t) L 1 mH SOLUTION 7.28 R - vo(t) 100 + At the heart of this circuit is a single-pole, double-throw switch—that is, a single switch (single-pole) with two electrically connected positions (double-throw), 1 and 2. As shown in Fig. 7.27a, when in position 1, the inductor current grows linearly in accordance with the equation i(t) = T1 1 V dt L 30 in Then the switch moves from position 1 to position 2 at time T1 . The peak inductor current is ip(t) = Vin T1 L While at position 1, the resistor is isolated electrically, and therefore its voltage is zero. irwin07_296-368hr.qxd 28-07-2010 11:34 Page 335 SECTION 7.4 Vin ± 5V – i(t) L R - Vin ± 5V – vo(t) 100 + 1 mH L i(t) 1 mH (a) A P P L I C AT I O N E X A M P L E S R 335 - vo(t) 100 + (b) Figure 7.27 (a) Pulse generator with switch in position 1. Inductor is energized. (b) Switch in position 2. As energy is drained from the inductor, the voltage and current decay toward zero. At time t 7 T1 , when the switch is in position 2, as shown in Fig. 7.27b, the inductor current flows into the resistor producing the voltage vo At - T1 B = iAt - T1 BR t 7 T1 At this point, we know that the form of the voltage vo(t), in the time interval t 7 T1 , is vo(t) = Ke-(t - T1)兾 t 7 T1 And = L兾R. The initial value of vo(t) in the time interval t 7 T1 is K since at time T1 the exponential term is 1. According to the design specifications, this initial value is 500 and therefore K = 500. Since this voltage is created by the peak inductor current IP flowing in R, K = 500 = AVin T1 RB兾L = 5T1(100)兾10-3 and thus T1 is 1 ms and IP is 5 A. The equation for the voltage in the time interval t 7 T1 , or t 7 1 ms, is vo(t - 1 ms) = 500e-100,000(t - 1 ms) V At the end of the 2-ms period, that is, at t = 2 ms, the voltage is 500e-100 or essentially zero. The complete waveform for the voltage is shown in Fig. 7.28. It is instructive at this point to consider the ratings of the various components used in this pulse generator circuit. First, 500 V is a rather high voltage, and thus each component’s voltage rating should be at least 600 V in order to provide some safety margin. Second, the inductor’s peak current rating should be at least 6 A. Finally, at peak current, the power losses in the resistor are 2500 W! This resistor will have to be physically large to handle this power load without getting too hot. Fortunately, the resistor power is pulsed rather than continuous; thus, a lower power rated resistor will work fine, perhaps 500 W. In later chapters we will address the issue of power in much more detail. Figure 7.28 Output Voltage (V) 600 The output voltage of the pulse generator. 500 400 300 200 100 0 0 10 20 30 Time (ms) 40 50 irwin07_296-368hr.qxd 336 28-07-2010 CHAPTER 7 11:34 Page 336 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S APPLICATION EXAMPLE 7.14 A heart pacemaker circuit is shown in Fig. 7.29. An SCR (silicon-controlled rectifier) is a solid-state device that has two distinct modes of operation. When the voltage across the SCR is increasing but less than 5 V, the SCR behaves like an open circuit, as shown in Fig. 7.30a. Once the voltage across the SCR reaches 5 V, the device functions like a current source, as shown in Fig. 7.30b. This behavior will continue as long as the SCR voltage remains above 0.2 V. At this voltage, the SCR shuts off and again becomes an open circuit. Assume that at t=0, vC(t) is 0 V and the 1-F capacitor begins to charge toward the 6-V source voltage. Find the resistor value such that vC(t) will equal 5 V (the SCR firing voltage) at 1 s. At t=1 s, the SCR fires and begins discharging the capacitor. Find the time required for vC(t) to drop from 5 V to 0.2 V. Finally, plot vC(t) for the three cycles. R Figure 7.29 Heart pacemaker equivalent circuit. + vC(t) - C V=6 V 1 F SCR Figure 7.30 Equivalent circuits for an SCR. SCR I=50 A SCR (a) SOLUTION (b) For t<1 s, the equivalent circuit for the pacemaker is shown in Fig. 7.31. As indicated earlier, the capacitor voltage has the form vC(t) = 6 - 6e-t兾RC V A voltage of 0.2 V occurs at t1 = 0.034RC whereas a voltage of 5 V occurs at t2 = 1.792RC We desire that t2 - t1 = 1 s. Therefore, t2 - t1 = 1.758RC = 1 s and RC = 0.569 s R = 569 k⍀ R Figure 7.31 Pacemaker equivalent network during capacitor charge cycle. and + V=6 V C 1 F vC(t) - irwin07_296-368hr.qxd 28-07-2010 11:34 Page 337 SECTION 7.4 A P P L I C AT I O N E X A M P L E S R Figure 7.32 + C V=6 V 337 vC(t) 1 F I=50 A Pacemaker equivalent network during capacitor discharge cycle. - At t=1 s the SCR fires and the pacemaker is modeled by the circuit in Fig. 7.32. The form of the discharge waveform is v(t) = K1 + K2 e-(t - 1)兾RC The term (t-1) appears in the exponential to shift the function 1 s, since during that time the capacitor was charging. Just after the SCR fires at t = 1+ s, vC(t) is still 5 V, whereas at t=q, vC(t)=6-IR. Therefore, K1 + K2 = 5 and K1 = 6 - IR Our solution, then, is of the form vC(t) = 6 - IR + (IR - 1)e-(t - 1)兾RC Let T be the time beyond 1 s necessary for v(t) to drop to 0.2 V. We write vC(T + 1) = 6 - IR + (IR - 1)e-T兾RC = 0.2 Substituting for I, R, and C, we find T = 0.11 s The output waveform is shown in Fig. 7.33. v(t) (V) Figure 7.33 6 Heart pacemaker output voltage waveform. 4 2 t(s) 0 0 1 2 3 4 Consider the simple RL circuit, shown in Fig. 7.34, which forms the basis for essentially every dc power supply in the world. The switch opens at t = 0. VS and R have been chosen simply to create a 1-A current in the inductor prior to switching. Let us find the peak voltage across the inductor and across the switch. APPLICATION EXAMPLE 7.15 • irwin07_296-368hr.qxd 338 28-07-2010 CHAPTER 7 11:34 Page 338 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S SOLUTION We begin the analysis with an expression for the inductor current. At t = 0, the inductor current is 1 A. At t = q, the current is 0. The time constant is simply L兾R, but when the switch is open, R is infinite and the time constant is zero! As a result, the inductor current is R iL(t) = 1e-␣t A 1 L 100 H VS ± 1V – + 7.29 where ␣ is infinite. The resulting inductor voltage is vL(t) vL(t) = L - di L(t) = -␣e-␣t dt 7.30 iL(t) At t = 0, the peak inductor voltage is negative infinity! This voltage level is caused by the attempt to disrupt the inductor current instantaneously, driving di兾dt through the roof. Employing KVL, the peak switch voltage must be positive infinity (give or take the supply v t=0 switch voltage). This phenomenon is called inductive kick, and it is the nemesis of power supply designers. Given this situation, we naturally look for a way to reduce this excessive voltage and, more importantly, predict and control it. Let’s look at what we have and what we know. We Figure 7.34 have a transient voltage that grows very quickly without bound. We also have an initial curThe switched inductor rent in the inductor that must go somewhere. We know that capacitor voltages cannot network at the heart of change quickly and resistors consume energy. Therefore, let’s put an RC network around modern power supplies. the switch, as shown in Fig. 7.35, and examine the performance that results from this change. + Figure 7.35 R 1 Conversion of a switched inductor circuit to an RLC network in an attempt to control inductive kick. L 100 H VS ± 1V – + vL(t) iL(t) t=0 + C vswitch - R The addition of the RC network yields a series circuit. We need the characteristic equation of this series RLC network when the switch is open. From Eq. (7.15), we know that the characteristic equation for the series RLC circuit is s2 + 20 s + 20 = s2 + c R + 1 1 ds + = 0 L LC 7.31 To retain some switching speed, we will somewhat arbitrarily choose a critically damped system where = 1 and 0 = 106 rad兾s. This choice for 0 should allow the system to stabilize in a few microseconds. From Eq. (7.31) we can now write expressions for C and R. 20 = 1012 = 1 1 = -4 LC 10 C 20 = 2 * 106 = R + 1 R + 1 = L 10-4 7.32 Solving these equations yields the parameter values C = 10 nF and R = 199 ⍀. Now we can focus on the peak switch voltage. When the switch opens, the inductor current, set at 1 A by the dc source and the 1-⍀ resistor, flows through the RC circuit. Since the capacitor was previously discharged by the closed switch, its voltage cannot change immediately and irwin07_296-368hr.qxd 28-07-2010 11:34 Page 339 SECTION 7.4 A P P L I C AT I O N E X A M P L E S 339 its voltage remains zero for an instant. The resistor voltage is simply IL R where IL is the initial inductor current. Given our IL and R values, the resistor voltage just after opening the switch is 199 V. The switch voltage is then just the sum of the capacitor and resistor voltages (i.e., 199 V). This is a tremendous improvement over the first scenario! A plot of the switch voltage, shown in Fig. 7.36, clearly agrees with our analysis. This plot illustrates the effectiveness of the RC network in reducing the inductive kick generated by opening the switch. Note that the switch voltage is controlled at a 199-V peak value and the system is critically damped; that is, there is little or no overshoot, having stabilized in less than 5 s. Because of its importance, this R-C network is called a snubber and is the engineer’s solution of choice for controlling inductive kick. Figure 7.36 Output Voltage (V) 600 Plot of the switch voltage when the snubber circuit is employed to reduce inductive kick. 200 200 200 200 100 0 0 10 20 30 40 50 Time (ms) AC voltage waveform One of the most common and necessary subcircuits that appears in a wide variety of electronic systems—for example, stereos, TVs, radios, and computers—is a quality dc voltage source or power supply. The standard wall socket supplies an alternating current (ac) voltage waveform shown in Fig. 7.37a, and the conversion of this voltage to a desired dc level is done as illustrated in Fig. 7.37b. The ac waveform is converted to a quasi-dc voltage by an inexpensive ac–dc converter whose output contains remnants of the ac input and is unregulated. A higher quality dc output is created by a switching dc–dc converter. Of the several versions of dc–dc converters, we will focus on a topology called the boost converter, shown in Fig. 7.38. Let us develop an equation relating the output voltage to the switching characteristics. + AC–DC converter Vin - Time (a) (b) DC–DC switching power supply APPLICATION EXAMPLE 7.16 Figure 7.37 (a) The ac voltage waveform at a standard wall outlet and (b) a block diagram of a modern dc power supply. + Vo - • irwin07_296-368hr.qxd 340 28-07-2010 CHAPTER 7 11:34 Page 340 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Figure 7.38 + vL(t) - The boost converter with switch settings for time intervals (a) ton and (b) toff. S2 + iL(t) Vin C R S1 Vo - (a) + vL(t) - S2 + iL(t) Vin C R Vo - S1 (b) SOLUTION Consider the boost converter in Fig. 7.38a, where switch 1 (S1) is closed and S2 is open for a time interval ton . This isolates the inductor from the capacitor, creating two subcircuits that can be analyzed independently. Note that during ton the inductor current and stored energy are increasing while at the output node, the capacitor voltage discharges exponentially into the load. If the capacitor’s time constant (=RC) is large, then the output voltage will decrease slowly. Thus, during ton energy is stored in the inductor and the capacitor provides energy to the load. Next, we change both switch positions so that S1 is open and S2 is closed for a time interval t off , as seen in Fig. 7.38b. Since the inductor current cannot change instantaneously, current flows into the capacitor and the load, recharging the capacitor. During toff the energy that was added to the inductor during ton is used to recharge the capacitor and drive the load. When toff has elapsed, the cycle is repeated. Note that the energy added to the inductor during ton must go to the capacitor and load during toff ; otherwise, the inductor energy would increase to the point that the inductor would fail. This requires that the energy stored in the inductor must be the same at the end of each switching cycle. Recalling that the inductor energy is related to the current by w(t) = 1 2 Li (t) 2 we can state that the inductor current must also be the same at the end of each switching cycle, as shown in Fig. 7.39. The inductor current during ton and toff can be written as iL(t) = ton Vin 1 ton 1 vL(t) dt = Vin dt = c d ton + I0 L 30 L 30 L iL(t) = ton + toff ton + toff 1 1 vL(t) dt = AVin - Vo B dt = L 3ton L 3ton c Vin - Vo d toff + I0 L ton 6 t 6 toff 0 6 t 6 ton 7.33 where I0 is the initial current at the beginning of each switching cycle. If the inductor current is the same at the beginning and end of each switching cycle, then the integrals in Eq. (7.33) must sum to zero. Or Vin ton = AVo - Vin Btoff = AVo - Vin B AT - ton B irwin07_296-368hr.qxd 28-07-2010 11:34 Page 341 SECTION 7.4 A P P L I C AT I O N E X A M P L E S 341 Figure 7.39 iL(t) I0 Vin vL(t) 0 Vin-Vo ton toff T 2T 3T Figure 7.40 30 Effect of duty cycle on boost converter gain. 25 Vo / Vin Waveform sketches for the inductor voltage and current. 20 15 10 5 0 0 20 40 60 Duty cycle (percent) 80 100 where T is the period AT = ton + toff B. Solving for Vo yields Vo = Vin c T 1 d = Vin c d = T - ton AT - ton B兾T Vin c 1 A1 - ton兾TB d = Vin c 1 d 1 - D where D is the duty cycle AD = ton兾TB. Thus, by controlling the duty cycle, we control the output voltage. Since D is always a positive fraction, Vo is always bigger than Vin —thus the name boost converter. A plot of Vo兾Vin versus duty cycle is shown in Fig. 7.40. An experimental schematic for a railgun is shown in Fig. 7.41. With switch Sw-2 open, switch Sw-1 is closed and the power supply charges the capacitor bank to 10 kV. Then switch Sw-1 is opened. The railgun is fired by closing switch Sw-2. When the capacitor discharges, the current causes the foil at the end of the gun to explode, creating a hot plasma that accelerates down the tube. The voltage drop in vaporizing the foil is negligible, and, therefore, more than 95% of the energy remains available for accelerating the plasma. The current flow establishes a magnetic field, and the force on the plasma caused by the magnetic field, which is proportional to the square of the current at any instant of time, accelerates the plasma. A higher initial voltage will result in more acceleration. The circuit diagram for the discharge circuit is shown in Fig. 7.42. The resistance of the bus (a heavy conductor) includes the resistance of the switch. The resistance of the foil and resultant plasma is negligible; therefore, the current flowing between the upper and lower conductors is dependent on the remaining circuit components in the closed path, as specified in Fig. 7.41. APPLICATION EXAMPLE 7.17 • irwin07_296-368hr.qxd 342 28-07-2010 CHAPTER 7 11:34 Page 342 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Sw–1 Figure 7.41 Sw–2 Experimental schematic for a railgun. Charging resistor Upper conductor Power supply 10 kV 100 kW Capacitor bank 53.6 F Foil Insulator Lower conductor Sw-2 Figure 7.42 Rbus=12 m Railgun discharge circuit. Lbus=32 nH i(t) + 10 kV - C=53.6 F Rfoil The differential equation for the natural response of the current is d2i(t) dt + 2 i(t) Rbus di(t) + = 0 Lbus dt Lbus C Let us use the characteristic equation to describe the current waveform. SOLUTION Using the circuit values, the characteristic equation is s2 + 37.5 * 104s + 58.3 * 1010 = 0 and the roots of the equation are s1 , s2 = (-18.75 ; j74) * 104 and hence the network is underdamped. The roots of the characteristic equation illustrate that the damped resonant frequency is d = 740 krad兾s Therefore, fd = 118 kHz and the period of the waveform is T = 1 = 8.5 s fd An actual plot of the current is shown in Fig. 7.43, and this plot verifies that the period of the damped response is indeed 8.5 s. Figure 7.43 300 Current (kA) Load current with a capacitor bank charged to 10 kV. 200 100 0 –100 –200 0 10 20 Time (microsecond) 30 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 343 SECTION 7.5 343 DESIGN EXAMPLES 7.5 Design Examples We wish to design an efficient electric heater that operates from a 24-V dc source and creates heat by driving a 1-⍀ resistive heating element. For the temperature range of interest, the power absorbed by the heating element should be between 100 and 400 W. An experienced engineer has suggested that two quite different techniques be examined as possible solutions: a simple voltage divider and a switched inductor circuit. In the first case, the required network is shown in Fig. 7.44. The variable resistance element is called a rheostat. Potentiometers are variable resistors that are intended for low power (i.e., less than 1 W) operation. Rheostats, on the other hand, are devices used at much higher power levels. We know from previous work that the heating element voltage is Vo = VS Rhe Rhe + Radj 7.34 Po = V2o Rhe = V2S 2 Rhe ARhe + Radj B 7.35 By substituting the maximum and minimum values of the output power into Eq. (7.35), we can determine the range of resistance required for the rheostat: Radj, min = Radj, max (242)(1) V2S Rhe - Rhe = - 1 = 0.2 ⍀ C Po, max C 400 7.36 (242)(1) V2S Rhe = - Rhe = - 1 = 1.4 ⍀ C Po, min C 100 So, a 2-⍀ rheostat should work just fine. But what about the efficiency of our design? How much power is lost in the rheostat? The rheostat power can be expressed as Padj = AVS - Vo B Radj 2 = V2S SOLUTION Radj VS 24 V Changing Radj will change the voltage across, and the power dissipated by, the heating element. The power can be expressed as Radj ARhe + Radj B 2 7.37 We know from our studies of maximum power transfer that the value of Radj that causes maximum power loss, and thus the worst-case efficiency for the circuit, occurs when Radj = Rhe = 1 ⍀. Obviously, the resistances consume the same power, and the efficiency is only 50%. Now that we understand the capability of this voltage-divider technique, let’s explore the alternative solution. At this point, it would at least appear that the use of a switched inductor is a viable alternative since this element consumes no power. So, if we could set up a current in an inductor, switch it into the heating element, and repeat this operation fast enough, the heating element would respond to the average power delivered to it and maintain a constant temperature. Consider the circuit in Fig. 7.45, where the switch moves back and forth, energizing the inductor with current, and then directing that current to the heating element. Let’s examine this concept to determine its effectiveness. We begin by assuming that the inductor current • DESIGN EXAMPLE 7.18 ± – Rhe 1 + Vo - Figure 7.44 A simple circuit for varying the temperature of a heating element. irwin07_296-368hr.qxd 344 28-07-2010 CHAPTER 7 11:34 Page 344 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Figure 7.45 iL(t) A switched inductor solution to varying the heating element’s temperature. L Rhe VS ± 24 V – 1 pos. 1 - vo(t) + pos. 2 is zero and the switch has just moved to position 1. The inductor current will begin to grow linearly in accordance with the fundamental equation iL(t) = VS 1 1 vL(t) dt = VS dt = t L 3 L 3 L 7.38 Note that VS兾L is the slope of the linear growth. Since VS is set at 24 V, we can control the slope with our selection for L. The inductor current increases until the switch moves at time t = t1, at which point the peak current is Ipeak = VS t L 1 7.39 This inductor current will discharge exponentially through the heating element according to the equation iL At¿B = Ipeak e-t¿兾 7.40 where t¿ is zero when the switch moves to position 2 and = L兾Rhe . If the switch is maintained in position 2 for about 5 time constants, the inductor current will essentially reach zero and the switch can return to position 1 under the initial condition—zero inductor current. A sketch of the inductor current over a single switching cycle is shown in Fig. 7.46. Repeated switching cycles will transfer power to the heating element. If the switching period is much shorter than the element’s thermal time constant—a measure of how quickly the element heats up—then the element’s temperature will be determined by the average power. This is a concept we don’t understand at this point. However, we will present Average Power in Chapter 9, and this example provides at least some motivation for its examination. Nevertheless, we should recognize two things: the load current is just the exponential decaying portion of the inductor current, and the initial value of that exponential is Ipeak, as defined in Eq. (7.39). Increasing Ipeak will increase the element’s power and temperature. As Eq. (7.39) indicates, this is easily done by controlling t1 ! It is impossible to proceed with the design until we can accurately predict the average power at the load. Figure 7.46 Ipeak Inductor current A single switching cycle for the inductor-based solution. The value of Ipeak is directly proportional to t1. VS /L t1 Time irwin07_296-368hr.qxd 28-07-2010 11:34 Page 345 SECTION 7.5 DESIGN EXAMPLES 345 Returning to our original concern, have we improved the efficiency at all? Note that there are no power-consuming components in our new circuit other than the heating element itself. Therefore, ignoring resistance in the inductor and the switch, we find that our solution is 100% efficient! In actuality, efficiencies approaching 95% are attainable. This is a drastic improvement over the other alternative, which employs a rheostat. • Consider the circuit in Fig. 7.47a, where a dc power supply, typically fed from a wall outlet, is modeled as a dc voltage source in series with a resistor RS . The load draws a constant current and is modeled as a current source. We wish to design the simplest possible circuit that will isolate the load device from disturbances in the power supply voltage. In effect, our task is to improve the performance of the power supply at very little additional cost. A standard solution to this problem involves the use of a capacitor CD, as shown in Fig. 7.47b. The two voltage sources and the single-pole double-throw switch model the input disturbance sketched in Fig. 7.47c. Engineers call CD a decoupling capacitor since it decouples disturbances in the input voltage from the output voltage. In typical electronic circuits, we find liberal use of these decoupling capacitors. Thus, our task is to develop a design equation for CD in terms of RS , VS , Vo , ¢VS , ¢Vo , and t¿. Our result will be applicable to any scenario that can be modeled by the circuit in Fig. 7.47b. RS t=t' t=0 RS IL t=t' IL t=0 + VS ± – ± – VS+ΔVS DESIGN EXAMPLE 7.19 CD vo(t) VS ± – ± V +ΔV S S – - VS+ΔVS + vo(t) - VS Vo+ΔVo Vo t' 0 (a) (b) (c) Figure 7.47 (a) A simple dc circuit that models disturbances in the source voltage. (b) The use of a decoupling capacitor to reduce disturbances in the load voltage. (c) Definitions of the input and output voltage disturbances. SOLUTION The voltage across CD can be expressed in the standard form as vo(t) = K1 + K2 e -t兾 7.41 Equivalent circuits for t = 0 and t = q are shown in Fig. 7.48a and b, respectively. At these two time extremes we find vo(0) = K1 + K2 = VS - IL RS vo(q) = K1 = VS + ¢VS - IL RS 7.42 To determine the time constant’s equivalent resistance, we return to the circuit in Fig. 7.48b, reduce all independent sources to zero, and view the resulting circuit from the capacitor’s terminals. It is easy to see that the time constant is simply RS CD . Thus, vo(t) = VS + ¢VS - IL RS - ¢VS e-t兾RS CD 7.43 t irwin07_296-368hr.qxd CHAPTER 7 11:34 Page 346 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S RS RS IL + VS ± – VS-ILRS vo(t) CD - - IL + + VS+ VS ± – VS+VS-ILRS + CD vo(t) - - (a) (b) Figure 7.48 The circuit in Fig. 7.47b at t = 0 just after the switch has moved. (b) The same circuit at t = q . At exactly t = t¿, the output voltage is its original value, Vo(0) plus ¢Vo . Substituting this condition into Eq. (7.43) yields vo At¿B = VS - IL RS + ¢Vo = VS + ¢VS - IL RS - ¢VS e-t¿兾RS CD which can be reduced to the expression ¢VS - ¢Vo = ¢VS e-t¿兾RS CD 7.44 Notice that the value of CD depends not on the input and output voltages, but rather on the changes in those voltages! This makes Eq. (7.44) very versatile indeed. A simple algebraic manipulation of Eq. (7.44) yields the design equation for CD: CD = t¿ RS ln c 7.45 ¢VS d ¢VS - ¢Vo Examining this expression, we see that CD is directly related to t¿ and inversely related to RS . If t¿ doubles or if RS is halved, then CD will double as well. This result is not very surprising. The dependence on the voltage changes is more complex. Let us isolate this term and express it as f = 1 1 -1 = = ¢VS ¢Vo 1 ln c d ln c d ln c 1 d 1 - ¢Vo兾¢VS ¢VS - ¢Vo ¢VS 7.46 Figure 7.49 shows a plot of this term versus the ratio ¢Vo兾¢VS . Note that for very small ¢Vo (i.e., a large degree of decoupling), this term is very large. Since this term is multiplied with t¿ in Eq. (7.45), we find that the price for excellent decoupling is a very large capacitance. Figure 7.49 A plot of the function f versus ¢Vo 兾¢VS. 100 Logarithmic term, f 346 28-07-2010 75 50 25 0 0 0.25 0.5 ΔVo /ΔVS 0.75 1 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 347 SECTION 7.5 DESIGN EXAMPLES Finally, as an example, consider the scenario in which VS is 5 V, RS is 20 ⍀, and the input disturbance is characterized by ¢VS = 1 V and t¿ = 0.5 ms. If the output changes are to be limited to only 0.2 V, the required capacitance would be CD = 112.0 F. Such a capacitor rated for operation at up to 16 V costs less than $0.20 and should be slightly smaller than a peanut M&M. This very simple, but very important, application demonstrates how an engineer can apply his or her basic circuit analysis skills to attack and describe a practical application in such a way that the result is broadly applicable. Remember, the key to this entire exercise is the creation of a circuit model for the decoupling scenario. The network in Fig. 7.50 models an automobile ignition system. The voltage source represents the standard 12-V battery. The inductor is the ignition coil, which is magnetically coupled to the starter (not shown). The inductor’s internal resistance is modeled by the resistor, and the switch is the keyed ignition switch. Initially, the switch connects the ignition circuitry to the battery, and thus the capacitor is charged to 12 V. To start the motor, we close the switch, thereby discharging the capacitor through the inductor. Assuming that optimum starter operation requires an overdamped response for iL(t) that reaches at least 1 A within 100 ms after switching and remains above 1 A for between 1 and 1.5 s, let us find a value for the capacitor that will produce such a current waveform. In addition, let us plot the response, including the time interval just prior to moving the switch, and verify our design. Switch t=0 VS=12 V Figure 7.50 vC (t) + C DESIGN EXAMPLE 7.20 iL(t) Circuit model for ignition system. L=200 mH R=4 Before the switch is moved at t=0, the capacitor looks like an open circuit, and the inductor acts like a short circuit. Thus, iL A0- B = iL A0+ B = 0 A and vC A0- B = vC A0+ B = 12 V After switching, the circuit is a series RLC unforced network described by the characteristic equation s2 + R 1 s + = 0 L LC with roots at s=–s1 and –s2 . The characteristic equation is of the form As + s1 B As + s2 B = s2 + As1 + s2 Bs + s1 s2 = 0 Comparing the two expressions, we see that and R = s1 + s2 = 20 L 1 = s1 s2 LC SOLUTION 347 irwin07_296-368hr.qxd 348 28-07-2010 CHAPTER 7 11:34 Page 348 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S Since the network must be overdamped, the inductor current is of the form iL(t) = K1 e-s1 t + K2 e-s2 t Just after switching, iL A0+ B = K1 + K2 = 0 or K2 = -K1 Also, at t = 0+, the inductor voltage equals the capacitor voltage because iL=0 and therefore iL R=0. Thus, we can write vL A0+ B = L di L A0+ B dt 1 - s1 K1 + s2 K1 = 12 L or K1 = 60 s2 - s1 At this point, let us arbitrarily choose s1=3 and s2=17, which satisfies the condition s1+s2=20, and furthermore, K1 = C = Hence, iL(t) is 60 60 = = 4.29 s2 - s1 14 1 1 = 98 mF = Ls1 s2 (0.2)(3)(17) iL(t) = 4.29C e-3t - e-17t D A Figure 7.51a shows a plot of iL(t). At 100 ms, the current has increased to 2.39 A, which meets the initial magnitude specifications. However, 1 second later, at t=1.1 s, iL(t) has fallen to only 0.16 A—well below the magnitude-over-time requirement. Simply put, the current falls too quickly. To make an informed estimate for s1 and s2 , let us investigate the effect the roots exhibit on the current waveform when s2>s1 . Since s2>s1 , the exponential associated with s2 will decay to zero faster than that associated with s1 . This causes iL(t) to rise—the larger the value of s2 , the faster the rise. After 5A1兾s2 B seconds have elapsed, the exponential associated with s2 is approximately zero and iL(t) decreases exponentially with a time constant of =1/s1 . Thus, to slow the fall of iL(t) iL(t) 3.0 A 3.0 A 2.0 A 2.0 A 1.0 A 1.0 A 0.0 A –0.0 s 0.5 s 1.0 s 1.5 s Time 2.0 s 2.5 s 2.9 s (a) Figure 7.51 Ignition current as a function of time. 0.0 A –0.0 s 0.5 s 1.0 s 1.5 s 2.0 s 2.5 s 3.0 s 3.5 s 4.0 s Time (b) irwin07_296-368hr.qxd 28-07-2010 11:34 Page 349 SECTION 7.5 DESIGN EXAMPLES 349 iL(t), we should reduce s1 . Hence, let us choose s1=1. Since s1+s2 must equal 20, s2=19. Under these conditions C = 1 1 = 263 mF = Ls1 s2 (0.2)(1)(19) and K1 = Thus, the current is 60 60 = = 3.33 s2 - s1 18 iL(t) = 3.33 C e-t - e-19t D A which is shown in Fig. 7.51b. At 100 ms the current is 2.52 A. Also, at t=1.1 s, the current is 1.11 A—above the 1-A requirement. Therefore, the choice of C=263 mF meets all starter specifications. DESIGN EXAMPLE 7.21 A defibrillator is a device that is used to stop heart fibrillations—erratic uncoordinated quivering of the heart muscle fibers—by delivering an electric shock to the heart. The Lown defibrillator was developed by Dr. Bernard Lown in 1962. Its key feature, shown in Fig. 7.52a, is its voltage waveform. A simplified circuit diagram that is capable of producing the Lown waveform is shown in Fig. 7.52b. Let us find the necessary values for the inductor and capacitor. vo(t) (V) L 3000 i(t) t=0 VS 6000 V 0 5 + R=50 patient C vo(t) - 10 Time (milliseconds) (a) (b) Figure 7.52 Lown defibrillator waveform and simplified circuit. Reprinted with permission from John Wiley & Sons, Inc., Introduction to Biomedical Equipment Technology. Since the Lown waveform is oscillatory in nature, we know that the circuit is underdamped ( 6 1) and the voltage applied to the patient is of the form vo(t) = K1 e-o t sin[t] where o = R 2L = o 21 - 2 and o = 1 1LC SOLUTION • irwin07_296-368hr.qxd 350 28-07-2010 CHAPTER 7 11:34 Page 350 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S for the series RLC circuit. From Fig. 7.52a, we see that the period of the sine function is T = 10 ms Thus, we have one expression involving o and = 2 = o 21 - 2 = 200 rad兾s T A second expression can be obtained by solving for the ratio of vo(t) at t = T兾4 to that at t = 3T兾4. At these two instants of time, the sine function is equal to +1 and -1, respectively. Using values from Fig. 7.52a, we can write vo(t兾4) -vo(3T兾4) = K1 e-o(T兾4) K1 e-o(3T兾4) = eo(T兾2) L 3000 = 12 250 or o = 497.0 Given R = 50 ⍀, the necessary inductor value is L = 50.3 mH Using our expression for , 2 = (200)2 = 2o - (o)2 or = (200)2 = 1 - (497.0)2 LC Solving for the capacitor value, we find C = 31.0 F Let us verify our design using the circuit shown in Fig. 7.53a. The output voltage plot shown in Fig. 7.53b matches the Lown waveform in Fig. 7.52a; thus, we can consider the design to be a success. While this solution is viable, it is not the only one. Like many design problems, there are often various ways to satisfy the design specifications. Figure 7.53 vo(t) Circuit and output plot for the Lown defibrillator. 4.0 kV L=50.3 mH 3.0 kV i(t) + vC(t)= 5000 V - + R=50 C=31 F vo(t) - 2.0 kV 1.0 kV 0 kV –1.0 kV 0s 2 ms 4 ms 6 ms Time (a) (b) 8 ms 10 ms irwin07_296-368hr.qxd 28-07-2010 11:34 Page 351 PROBLEMS 351 SUMMARY • First-Order Circuits ■ ■ ■ An RC or RL transient circuit is said to be first order if it contains only a single capacitor or single inductor. The voltage or current anywhere in the network can be obtained by solving a first-order differential equation. Second-Order Circuits ■ The form of a first-order differential equation with a constant forcing function is x(t) dx(t) dt + 2 ■ • The time constant for an RC circuit is RTh C and for an RL circuit is L/RTh , where RTh is the Thévenin equivalent resistance looking into the circuit at the terminals of the storage element (i.e., capacitor or inductor). The two approaches proposed for solving first-order transient circuits are the differential equation approach and the step-by-step method. In the former case, the differential equation that describes the dynamic behavior of the circuit is solved to determine the desired solution. In the latter case, the initial conditions and the steady-state value of the voltage across the capacitor or current in the inductor are used in conjunction with the circuit’s time constant and the known form of the desired variable to obtain a solution. + 20 ■ The characteristic equation for a second-order circuit is s2 + 20 s + 20 = 0, where is the damping ratio and 0 is the undamped natural frequency. ■ If the two roots of the characteristic equation are where A is referred to as the steady-state solution and is called the time constant. The function e-t兾 decays to a value that is less than 1% of its initial value after a period of 5. Therefore, the time constant, , determines the time required for the circuit to reach steady state. dx(t) + 20 x(t) = f(t) dt dt where f(t) is the network forcing function. = A x(t) = A + K2 e-t兾 ■ The voltage or current in an RLC transient circuit can be described by a constant coefficient differential equation of the form d2x(t) and the solution is ■ The response of a first-order transient circuit to an input pulse can be obtained by treating the pulse as a combination of two step-function inputs. ■ ■ ■ ■ real and unequal, then >1 and the network response is overdamped real and equal, then =1 and the network response is critically damped complex conjugates, then <1 and the network response is underdamped The three types of damping together with the corresponding network response are as follows: 1. Overdamped: 2 2 x(t) = K1 e-A0 - 0 2 - 1Bt + K2 e-A0 + 0 2 - 1Bt 2. Critically damped: x(t) = B1 e-0 t + B2 te-0 t 3. Underdamped: x(t) = e-t AA 1 cos d t + A 2 sin d tB, where = 0 and d = 0 21 - 2 ■ Two initial conditions are required to derive the two unknown coefficients in the network response equations. PROBLEMS 7.1 Use the differential equation approach to find i(t) for t 7 0 in the network in Fig. P7.1. 7.2 Use the differential equation approach to find io(t) for t 7 0 in the network in Fig. P7.2. t=0 4 k 6 12 V ± – 1 k t=0 i(t) 2H 6 2 k 4 mA 300 F 2 k io(t) Figure P7.1 Figure P7.2 irwin07_296-368hr.qxd 352 28-07-2010 CHAPTER 7 11:34 Page 352 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.3 Use the differential equation approach to find vo(t) 7.7 Use the differential equation approach to find io(t) for for t 7 0 in the circuit in Fig. P7.3 and plot the response including the time interval just prior to switch action. 12 k t=0 t 7 0 in the circuit in Fig. P7.7 and plot the response including the time interval just prior to opening the switch. 2 k 50 F t=0 12 V + + 6V ± – 100 F 4 k vo(t) 3 k 4 k - 4 k 12 k 12 k 8 k io(t) Figure P7.3 Figure P7.7 7.4 Use the differential equation approach to find vC(t) for 7.8 Use the differential equation approach to find io(t) for t 7 0 in the circuit in Fig. P7.4. t 7 0 in the circuit in Fig. P7.8 and plot the response including the time interval just prior to closing the switch. t=0 6 k 12 k 200 F 6 k + vC(t) 3 k ± – 100 F ± – 6V - t=0 12 k 12 V 6 k io(t) Figure P7.8 Figure P7.4 7.5 Use the differential equation approach to find vC(t) for 7.9 Use the differential equation approach to find vo(t) for t 7 0 in the circuit in Fig. P7.9 and plot the response including the time interval just prior to opening the switch. t 7 0 in the circuit in Fig. P7.5. 9 k 100 F 4 k 3 k 6V ± – + vC(t) t=0 100 F + 6 k - t=0 12 V ± – 6 k 6 k vo(t) - Figure P7.5 Figure P7.9 7.6 Use the differential equation approach to find vC(t) for t 7 0 in the circuit in Fig. P7.6 and plot the response including the time interval just prior to closing the switch. 2 k 4 k 7.10 In the network in Fig. P7.10, find io(t) for t 7 0 using the differential equation approach. 6 4 k io(t) + 6V ± – vC(t) t=0 Figure P7.6 t=0 100 F 2 k 2H - Figure P7.10 4 2A 12 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 353 PROBLEMS 7.11 Use the differential equation approach to find iL(t) for 7.14 Using the differential equation approach, find io(t) for t 7 0 in the circuit in Fig. P7.11 and plot the response, including the time interval just prior to opening the switch. 6 6 + 6 iL(t) t=0 6V 12 t 7 0 in the circuit in Fig. P7.14 and plot the response, including the time interval just prior to opening the switch. 2H + 353 6V 2H 3 12 V ± – 12 4 Figure P7.11 t=0 io(t) Figure P7.14 7.12 Use the differential equation approach to find vo(t) for t 7 0 in the circuit in Fig. P7.12 and plot the response, including the time interval just prior to opening the switch. 7.15 Use the step-by-step technique to find io(t) for t 7 0 in the network in Fig. P7.15. 100 F 2 k 4 k io(t) 6V 12 k + ± – ± – 12 V + 200 F 6 k t=0 12 V 8 k t=0 Figure P7.15 vo(t) 7.16 Use the step-by-step method to find io(t) for t 7 0 in the 4 k circuit in Fig. P7.16. 2 k Figure P7.12 Figure P7.16 7.13 Use the differential equation approach to find i(t) for 5 k t=0 2 k 7.17 Find vo(t) for t>0 in the network in Fig. P7.17. t=0 1 k i(t) 5 mA 1 k 4 k 2 k 12 V t=0 io(t) t 7 0 in the circuit in Fig. P7.13 and plot the response, including the time interval just prior to opening the switch. ± – 2H 6 k 1 k 4 k 1 mH 6 mA 4 k + 4 k 100 F vo(t) - Figure P7.13 Figure P7.17 irwin07_296-368hr.qxd 354 28-07-2010 CHAPTER 7 11:34 Page 354 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.18 Find vo(t) for t>0 in the network in Fig. P7.18. 7.23 Find vo(t) for t>0 in the circuit in Fig. P7.23. t=0 3 k 6 k 6 k 6 k 2 k 50 F t=0 12 mA 6 k 4 k vo(t) 12 V + 3 k 8 k 2 k + - 100 F vo(t) 1 k - Figure P7.23 Figure P7.18 7.19 Use the step-by-step method to find io(t) for t 7 0 in the circuit in Fig. P7.19. 7.24 Find vo(t) for t>0 in the network in Fig. P7.24. t=0 6 ± – 2H io(t) 6V + - 6 12 k 6 k + 8 k 50 F 24 V vo(t) 4 k - t=0 Figure P7.24 Figure P7.19 7.25 Use the step-by-step technique to find io(t) for t 7 0 in the network in Fig. P7.25. 7.20 Use the step-by-step method to find io(t) for t 7 0 in the circuit in Fig. P7.20. ±– 4 k 12 V 4 k t=0 4 k 200 F 12 k 12 mA 6 k 200 F 4 k t=0 io(t) 12 k Figure P7.25 io(t) 7.26 Use the step-by-step method to find vo(t) for t 7 0 in Figure P7.20 the network in Fig. P7.26. t=0 7.21 Find vo(t) for t>0 in the circuit in Fig. P7.21. 80 F 6 k + t=0 4 mA 6 k 6 k + 6 k 12 mA 100 F vo(t) - Figure P7.21 7.27 Find io(t) for t>0 in the circuit in Fig. P7.27. 9 k 2 k 12 mA 60 F 4 k 6 k t=0 io(t) t=0 2 k + vo(t) 4 VA + - - Figure P7.22 vo(t) - Figure P7.26 7.22 Find vo(t) for t>0 in the circuit in Fig. P7.22. 3 k 6 k Figure P7.27 1 k 150 F 1 k + - 6V irwin07_296-368hr.qxd 28-07-2010 11:34 Page 355 PROBLEMS 7.28 Find io(t) for t>0 in the circuit in Fig. P7.28. 355 7.32 Use the step-by-step method to find vo(t) for t 7 0 in the network in Fig. P7.32. 6V –+ 1 —H 3 t=0 2 k + 4 t=0 2 k 100 F 2 k 2 io(t) ± – Figure P7.28 12 V 7.29 Find io(t) for t>0 in the circuit in Fig. P7.29. – ± 12 V 4 k vo(t) 2 - Figure P7.32 4 mA 8 k 120 F 4 k t=0 7.33 Find vc(t) for t 7 0 in the network in Fig. P7.33 using the step-by-step method. it (t) 2 k 2 k Figure P7.29 t=0 7.30 Find io(t) for t>0 in the network in Fig. P7.30. + 50 F vC(t) 2 k 100 F - ± – 6 k 6 k + - t=0 6V 12 V io(t) 6 k Figure P7.33 7.34 Find vo(t) for t 7 0 in the network in Fig. P7.34 using Figure P7.30 the step-by-step method. 7.31 Find io(t) for t>0 in the circuit in Fig. P7.31. t=0 2 k 150 F 50 F io(t) ± – 2 k 24 V + 2 k 10 k 4 k 6 mA t=0 4 k vo(t) - Figure P7.31 Figure P7.34 4 k ± – 12 V irwin07_296-368hr.qxd 356 28-07-2010 CHAPTER 7 11:34 Page 356 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.35 Find vo(t) for t 7 0 in the circuit in Fig. P7.35 using the 7.38 Find vo(t) for t 7 0 in the circuit in Fig. P7.38 step-by-step method. using the step-by-step method. t=0 4 k 3 6 vo(t) 4 k + + 50 F 8 k 4 12 V 6 6A + vo(t) t=0 2H Figure P7.38 - 7.39 The switch in the circuit in Fig. P7.39 is opened at t=0. Find i(t) for t>0. Figure P7.35 0.5 H t=0 4 7.36 Use the step-by-step method to find io(t) for t 7 0 in the circuit in Fig. P7.36. 4 k 24 V 4 k 200 F 6V ± – t=0 4 k i(t) 3 2 k + – 2 – + io(t) 9V Figure P7.36 Figure P7.39 7.40 The switch in the circuit in Fig. P7.40 is moved at t=0. Find io(t) for t>0. 7.37 Find vo(t) for t 7 0 in the network in Fig. P7.37 using the step-by-step technique. 4 k 1 k 200 F 4 k + 24 V ± – 4 k t=0 2 k vo(t) - 18 V 100 F + t=0 4 k 3 k io(t) 4 k Figure P7.40 Figure P7.37 7.41 The switch in Fig. P7.41 closes at t=0. Find vo(t) for t>0. t=0 9 k + 6 k 6V + - 120 F 3 k 4 mA 4 k vo(t) - Figure P7.41 4 k irwin07_296-368hr3.qxd 2-09-2010 12:13 Page 357 357 PROBLEMS 7.42 The switch in the circuit in Fig. P7.42 has been closed for a long time and is opened at t=0. Find i(t) for t>0. 1 k⍀ 7.46 Find io(t) for t 7 0 in the network in Fig. P7.46 using the step-by-step method. 12 V 1 k⍀ + 2 k⍀ t=0 + - 30 V 200 F 0.3 H i(t) + - 10 V t=0 2 k⍀ 2 k⍀ 2 k⍀ 2 k⍀ 4 k⍀ 2 k⍀ io(t) Figure P7.42 7.43 The switch in the circuit in Fig. P7.43 has been closed for a long time and is moved at t=0. Find io(t) for t>0. Figure P7.46 7.47 Use the step-by-step method to find io(t) for t 7 0 in the t=0 5 k⍀ network in Fig. P7.47. io(t) - 10 V + 5 mA 5 k⍀ 6 k⍀ 2 k⍀ 6 k⍀ 2/3 H 6 k⍀ 5 k⍀ + - io(t) t=0 12 V 50 F Figure P7.43 7.44 The switch in the circuit in Fig. P7.44 has been in position A for a long time and is moved to position B at t=0. Calculate vo(t) for t>0. Figure P7.47 7.48 Find vo(t) for t>0 in the network in Fig. P7.48. t=0 A 5 k⍀ 3 k⍀ B + 24 V + - 1 k⍀ vo(t) 10 F + - + - t=0 8 k⍀ 60 F 6V 6 k⍀ 12 V + 80 F - vo(t) 4 k⍀ - Figure P7.44 Figure P7.48 7.45 The switch in the circuit in Fig. P7.45 has been closed for a long time and is opened at t=0. Determine vC(t) for t>0. 5 k⍀ 5 k⍀ 7.49 Find vo(t) for t>0 in the network in Fig. P7.49. 3 k⍀ t=0 9 k⍀ t=0 30 V 2 k⍀ + + - Figure P7.45 + 10 k⍀ + 18 V 20 F vC (t) 4V 6 k⍀ - + 8 k⍀ 200 F 4 k⍀ vo(t) - Figure P7.49 irwin07_296-368hr.qxd 358 28-07-2010 CHAPTER 7 11:34 Page 358 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.50 Find vo(t) for t>0 in the network in fig. P7.50. 7.54 Find vo(t) for t>0 in the network in Fig. P7.54. 12 V + t=0 6 k 3 k 4 k 12 k 2 k 6 k 3 k 3 k + 12 V + - 160 F 3 k vo(t) 200 F + 6V - + - 2 k vo(t) 1 k - Figure P7.50 t=0 7.51 Determine vo(t) for t>0 in the network in Fig. P7.51 Figure P7.54 7.55 Use the step-by-step technique to find vo(t) for t 7 0 in t=0 6 k 5 k the circuit in Fig. P7.55. + - + 6V 50 F 6 k 3 k + 12 V + - 200 F - t=0 6 k + - Figure P7.51 7.52 Find vo(t) for t>0 in the circuit in Fig. P7.52. 4 k 7.56 Use the step-by-step technique to find io(t) for t 7 0 in + - 3 k the network in Fig. P7.56. 6V 4 mA 100 F + 12 V + - 12 V Figure P7.55 1 k 2 k - 6 k vo(t) 1 k vo(t) 6 k t=0 100 F 4 k 2 k vo(t) 4 k 2 k 2 k 12 k 2 k - t=0 ± – 4V io(t) Figure P7.52 Figure P7.56 7.53 Find vo(t) for t>0 in the network in Fig. P7.53. + 150 F 6 k 3 k vo(t) - t=0 6 mA 7.57 Find vo(t) for t 7 0 in the network in Fig. P7.57 using the step-by-step technique. + 2 2H 3 12 V 12 k 2 ± – 6 4 k t=0 Figure P7.53 vo(t) Figure P7.57 - irwin07_296-368hr2.qxd 2-08-2010 16:43 Page 359 359 PROBLEMS 7.62 Find vo(t) for t>0 in the circuit in Fig. P7.62. 7.58 Find io(t) for t>0 in the network in Fig. P7.58. t=0 1 k⍀ 4 k⍀ 150 F 6V t=0 3 k⍀ + - 1 k⍀ 4 mA 1 k⍀ + - 6V 2 k⍀ + 1 k⍀ io(t) 200 F 1 k⍀ 1 k⍀ Figure P7.58 vo(t) - 7.59 Find vc(t) for t 7 0 in the circuit in Fig. P7.59 using Figure P7.62 the step-by-step method. 5⍀ 7.63 Find io(t) for t>0 in the network in Fig. P7.63. t=0 io(t) 5⍀ 20 ⍀ 6⍀ t=0 0.5 F 6A - vC(t) 12 V + 10 ⍀ 3⍀ + – 12 ⍀ + 0.2 H 10 ⍀ Figure P7.63 5⍀ 7.64 Find vo(t) for t>0 in the circuit in Fig. P7.64. Figure P7.59 6⍀ 7.60 Find i(t) for t 7 0 in the circuit in Fig. P7.60 using the step-by-step method. 6⍀ t=0 5⍀ t=0 12 V 2⍀ + 0.5 H ± – i(t) 2⍀ –30 V + t=0 2⍀ 30 V ± – 6⍀ 0.1 H vo(t) 6⍀ - Figure P7.64 Figure P7.60 7.61 Find i(t) for t 7 0 in the circuit in Fig. P7.61 using the step-by-step method. 7.65 Find io(t) for t>0 in the circuit in Fig. P7.65. 2H 12 ⍀ 10 ⍀ 105 V 15 ⍀ t=0 1H ± – 6V + 6⍀ i(t) 6⍀ 6⍀ 4⍀ t=0 2⍀ io(t) Figure P7.61 Figure P7.65 6V irwin07_296-368hr.qxd 360 28-07-2010 CHAPTER 7 11:34 Page 360 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.66 Find io(t) for t>0 in the network in Fig. P7.66. 3 + – 12 V 2 1H t=0 6 4 4A io(t) Figure P7.66 7.67 Find io(t) for t>0 in the circuit in Fig. P7.67. 2H 6 6V + – t=0 4 12 2 2A io(t) Figure P7.67 7.68 Find vo(t) for t>0 in the network in Fig. P7.68. 1H + 6 3A vo(t) 2 t=0 3 + – 2 24 A - Figure P7.68 7.71 Use the step-by-step method to find vo(t) for t 7 0 in 7.69 Find io(t) for t>0 in the circuit in Fig.P7.69. the network in Fig. P7.71. io(t) 2 6 k 6 12 V 6 vo(t) 4 k t=0 3 1H 6 F 6 12 V + – + 2 k t=0 ± – 6 F 6 F - Figure P7.71 Figure P7.69 7.70 Use the step-by-step method to find io(t) for t 7 0 in the network in Fig. P7.70. 7.72 Find iL(t) for t 7 0 in the circuit in Fig. P7.72 using the step-by-step method. 4 t=0 t=0 6 12 V ± – 4 12 io(t) 10 V ± – 3H 3A 2H 12 + 4A v1(t) - 2 2H ±– 5v1(t) Figure P7.70 Figure P7.72 iL(t) irwin07_296-368hr.qxd 28-07-2010 11:34 Page 361 PROBLEMS 7.73 Find io(t) for t 7 0 in the network in Fig. P7.73 using the step-by-step method. iA 2 k 7.76 The current source in the network in Fig. P7.76a is defined in Fig. P7.76b. The initial voltage across the capacitor must be zero. (Why?) Determine the current io(t) for t 7 0. t=0 2 k io(t) 10 mA 2 k 2 2 mH 361 2iA 2 k i(t) 2F 2 2 io(t) Figure P7.73 (a) i(t) (A) 6 7.74 Use the step-by-step technique to find vo(t) for t 7 0 in the network in Fig. P7.74. 24 V 3 k ±– 6V 200 F 3 k 2 k 0 4.5 (b) + – ± vo(t) t=0 - t (s) 6 k 2000 iA ± – iA Figure P7.76 7.77 The voltage v(t) shown in Fig. P7.77a is given by the graph shown in Fig. P7.77b. If iL(0) = 0, answer the following questions: (a) How much energy is stored in the inductor at t = 3 s? Figure P7.74 (b) How much power is supplied by the source at t = 4 s? (c) What is i(t = 6 s)? (d) How much power is absorbed 7.75 Determine the equation for the voltage vo(t) for t 7 0 in by the inductor at t = 3 s? Fig. P7.75a when subjected to the input pulse shown in Fig. P7.75b. 3 k 2 k + v(t) ± – i(t) vo(t) 200 F 6 k v(t) ± – 2 2H iL(t) (a) (a) v(t) v(t) (V) 10 V 12 5 t(s) 2 0 1 t (s) –10 V (b) Figure P7.75 (b) Figure P7.77 irwin07_296-368hr.qxd 362 28-07-2010 CHAPTER 7 11:34 Page 362 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.78 Find the output voltage vo(t) in the network 7.82 The switch in the circuit in Fig. P7.82 is closed at t = 0. If i1(0-) = 2 A, determine i2(0+), vR(0+), and i1(t = q) . in Fig. P7.78 if the input voltage is vi(t) = 5(u(t) - u(t - 0.05))V. t=0 i1(t) vi(t) ± – + + 1 F vR(t) 2H vo(t) 100 k i2(t) 4 3H - - Figure P7.82 Figure P7.78 7.83 The switch in the circuit in Fig. P7.83 has been closed for a long time and is opened at t = 0. If vC(t) = 20 - 8e-0.05t V, find R1, R2, and C. 7.79 Given that vC1(0-) = -10 V and vC2(0-) = 20 V in the circuit in Fig. P7.79, find i(0+). + R2 t=0 4 + i(t) vC1(t) 2F 4F - + vC(t) C R1 2A vC2(t) t=0 - - Figure P7.79 Figure P7.83 7.84 Given that i(t) = 2.5 + 1.5e-4t A for t 7 0 in the 7.80 In the network in Fig. P7.80, find i(t) for t 7 0. If circuit in Fig. P7.84, find R1, R2, and L. t=0 vC1(0-) = -10 V, calculate vC2(0-). i(t) i(t) vC1(t) + - 0.3 F ± – 10e–5tu(t) V R1 + ± – vC2(t) 0.6 F - R2 L 20 V Figure P7.84 7.85 Given that i(t) = 13.33e-t - 8.33e-0.5tA for t 7 0 in Figure P7.80 the network in Fig. P7.85, find the following: (a) vc(0), 7.81 In the circuit in Fig. P7.81, vR(t) = 100e-400t V for t 6 0. Find vR(t) for t 7 0. (b) vc(t = 1 s), (c) the capacitance C. t=0 + 25 100 vR(t) 4 L ± – 40e–tu(t) V - Figure P7.81 v (t) + C - i(t) Figure P7.85 C 2 irwin07_296-368hr.qxd 28-07-2010 11:34 Page 363 363 PROBLEMS 7.86 The differential equation that describes the current io(t) 7.92 For the underdamped circuit shown in Fig. P7.92, determine the voltage v(t) if the initial conditions on the storage elements are iL(0) = 1 A and vC(0) = 10 V. in a network is 2 d io(t) dt 2 + 6 dio(t) dt + 4io(t) = 0 + iL(0) Find (a) the characteristic equation of the network, v(t) 2H (b) the network’s natural frequencies dt + 4 dv1(t) dt + 5v1(t) = 0 1 —F 40 vC (0) - Figure P7.92 7.87 The voltage v1(t) in a network is defined by the equation 2 5 - (c) the expression for io(t). d 2v1(t) + 7.93 Find vC(t) for t 7 0 in the circuit in Fig. P7.93. t=0 Find (a) the characteristic equation of the network (b) the circuit’s natural frequencies 0.04 F 8 1A (c) the expression for v1(t) . + iL(t) vC(t) 1H - 7.88 The voltage v1(t) in a network is defined by the equation d2v1(t) dt 2 + 2c dv1(t) dt Figure P7.93 d + 5v1(t) = 0 7.94 Find vC(t) for t 7 0 in the circuit in Fig. P7.94 if Find vC(0) = 0. (a) the characteristic equation of the network. t=0 (b) the circuit’s natural frequencies. 1 k 100 mH (c) the expression for v1(t). + 7.89 A parallel RLC circuit contains a resistor R = 1 ⍀ and ± – 1 F 12 V vC(t) - an inductor L = 2H. Select the value of the capacitor so that the circuit is critically damped. Figure P7.94 7.90 A series RLC circuit contains a resistor of R = 2 ⍀ and a capacitor C = 1兾2 F. Select the value of the inductor so that the circuit is critically damped. 7.95 Find vo(t) for t 7 0 in the circuit in Fig. P7.95 and plot the response, including the time interval just prior to closing the switch. 2 7.91 In the critically damped circuit shown in Fig. P7.91, the t=0 initial conditions on the storage elements are iL(0) = 2 A and vC(0) = 5 V. Determine the voltage v(t). + vC (0) + 0.01 F - Figure P7.91 v(t) iL(0) 10 + 12 V ± – 4H - Figure P7.95 1 —F 400 1H 10 vo(t) - irwin07_296-368hr.qxd 364 28-07-2010 CHAPTER 7 11:34 Page 364 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.96 The switch in the circuit in Fig. P7.96 has been closed for 7.97 The switch in the circuit in Fig P7.97 has been closed for a a long time and is opened at t = 0. Find i(t) for t 7 0. long time and is opened at t = 0. Find i(t) for t 7 0. 6 t=0 5 1.5 5 t=0 5 20 V 1H – ± 12 V 0.04 F ± – 0.1 F ± – i(t) ± – i(t) 6V 10 V 1.25 H Figure P7.96 Figure P7.97 7.98 Find vo(t) for t 7 0 in the circuit in Fig. P7.98 and plot the response, including the time interval just prior to closing the switch. 1 —H 8 8 1 —H 8 t=0 + 1 —F 25 4 1.5 A 6 vo(t) 1 - Figure P7.98 7.99 In the circuit shown in Fig. P7.99, find v(t) 7 0. 7.101 Find vo(t) for t 7 0 in the circuit in Fig. P7.101 and plot the response, including the time interval just prior to moving the switch. 12 H — 5 t=0 t=0 2.5 mH 8 k + ± – 1 F — 12 2 8V v(t) 2 — nF 15 1 k 6 k ± – 6 k 7.100 Find vo(t) for t 7 0 in the circuit in Fig. P7.100 and plot the response, including the time interval just prior to moving the switch. t=0 plot the response, including the time interval just prior to moving the switch. 1 mH 1 k 0.05 nF vo(t) Figure P7.101 7.102 Find vo(t) for t 7 0 in the network in Fig. P7.102 and 10 k 4 k 3 mA - 12 V 2.5 mA 4 k + Figure P7.99 100 V ± – Figure P7.100 5 k t=0 + 10 k 10 F vo(t) - 100 mH Figure P7.102 2 k irwin07_296-368hr.qxd 28-07-2010 11:34 Page 365 PROBLEMS 7.103 Find vo(t) for t 7 0 in the circuit in Fig. P7.103 and plot the response, including the time interval just prior to moving the switch. the characteristic equation s2 + 4 * 107s + 3 * 1014 = 0 ± t=0 ± – 7.104 Design a parallel RLC circuit with R ⭌ 1 k⍀ that has 1 mH 2 k 12 V 365 1 k 6.25 nF vo(t) 7.105 Design a parallel RLC circuit with R ⭌ 1 k⍀ that has – the characteristic equation s 2 + 4 * 107s + 4 * 1014 = 0 Figure P7.103 Solve the remaining problems using computational methods. 7.106 Given the network in Fig. P7.106 to plot vo(t) over a 10-s interval starting at t = 0, using a 100-ms step size. 2 t=0 4 2H ± – 12 V 6 2A 1F + 2 vo(t) - Figure P7.106 7.107 Given the network in Fig. P7.107, plot vo(t) over a 10-s interval starting at t = 0, using a 100-ms step size. –± t=0 20 k 6V age shown in Fig. P7.108b, plot the voltage vo(t) over the interval 0 ⱕ t ⱕ 4 s, using a 20-ms step size. 12 V 100 k 10 k ± – + 10 mH 7.108 Given the network in Fig. P7.108a and the input volt- 5 k 50 F vin(t) ± – + 1M vo(t) 0.1 F 2 mH - vo(t) - (a) vin(t) V Figure P7.107 10 0 1 t(s) (b) Figure P7.108 irwin07_296-368hr.qxd 366 28-07-2010 CHAPTER 7 11:34 Page 366 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7.109 Given the network in Fig. P7.109a, and the input voltage shown in Fig. P7.109b, plot vo(t) over the interval 0 ⱕ t ⱕ 10 s, using a 100-ms step size. –± vin(t) 2 2 2 2A 1F 1 1H + vo(t) 4 - (a) vin(t) V 10 0 1 t (s) (b) Figure P7.109 7.110 Given the network in Fig. P7.110a and the input in Fig. P 7.110b, plot vo(t) in the interval 0 ⱕ t ⱕ 4 s, using a 20-ms step size. 10 k 5 k vs(t) ± – + 100 F 10 mH 10 k vo(t) - (a) vs(t) V 10 0 1 2 3 (b) Figure P7.110 4 t (s) irwin07_296-368hr.qxd 28-07-2010 11:34 Page 367 TYPICAL PROBLEMS FOUND ON THE FE EXAM • TYPICAL PROBLEMS FOUND ON THE FE EXAM 7PFE-1 In the circuit in Fig. 7PFE-1, the switch, which has been closed for a long time, opens at t = 0. Find the value of the capacitor voltage vc(t) at t = 2s. a. 0.936 V b. 0.756 V c. 0.264 V d. 0.462 V t=0 8 k 12 V 6 k + ± – vC(t) 6 k 100 F 6 k - Figure 7PFE-1 7PFE-2 In the network in Fig. 7PFE-2, the switch closes at t = 0. Find v0(t) at t = 1s. a. 5.62 V b. 1.57 V c. 4.25 V d. 3.79 V 12 k 4 k + t=0 ± – 12 V 100 F 12 k vo(t) - Figure 7PFE-2 7PFE-3 Assume that the switch in the network Fig. 7PFE-3 has been closed for some time. At t = 0 the switch opens. Determine the time required for the capacitor voltage to decay to one-half of its initially charged value. a. 0.416 s b. 0.625 s c. 0.235 s d. 0.143 s 12 k t=0 12 V ± – + vC(t) - Figure 7PFE-3 100 F 6 k 367 irwin07_296-368hr.qxd 368 28-07-2010 CHAPTER 7 11:34 Page 368 F I R S T- A N D S E C O N D - O R D E R T R A N S I E N T C I R C U I T S 7PFE-4 Find the inductor current iL(t) for t 7 0 in the circuit in Fig. 7PFE-4. a. iL(t) = 3 - 2e-t兾6 A, t 7 0 b. iL(t) = 1 + 2e-2t兾3 A, t 7 0 c. iL(t) = 6 - e-t兾6 A, t 7 0 d. iL(t) = 3 - e-2t兾3 A, t 7 0 t=0 2 iL(t) ± – 10 V 2 2 4H 1A Figure 7PFE-4 7PFE-5 Find the inductor current iL(t) for t 7 0 in the circuit in Fig. 7PFE-5. a. iL(t) = 1.4 + 0.4e-4t兾3 A, t 7 0 b. iL(t) = 1.2 + 0.4e-5t兾3 A, t 7 0 c. iL(t) = 0.4 + 0.2e-4t兾3 A, t 7 0 d. iL(t) = 2.4 + 0.6e-5t兾3 A, t 7 0 t=0 iL(t) 3H 3 12 V ± – Figure 7PFE-5 2 4 4 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 369 C H A PT E R AC STEADY-STATE ANALYSIS 8 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Understand the basic characteristics of sinusoidal functions ■ Be able to perform phasor and inverse phasor transformations and draw phasor diagrams ■ Know how to calculate impedance and admittance for our basic circuit elements: R, L, C ■ Be able to combine impedances and admittances in series and parallel ■ Be able to draw the frequency-domain circuit for a given circuit with a sinusoidal source ■ Know how to apply our circuit analysis techniques to frequency-domain circuits TebNad/iStockphoto S Smart Power Grids What makes a power grid smart? It takes a two-way communication between the electric power generator and the consumer. For providers, a smart grid gives rapid information tricity continues to increase, a smarter grid that provides electricity more efficiently also helps in the fight against global warming. This chapter introduces ac signals—alternating voltages and about power quality and alerts them when blackouts occur. For currents—which is the form of electricity that the power grid deliv- consumers, it promotes the efficient use of electricity by letting ers to the end user. Your background in differential equations, them know the time of day when power is more available and less needed in the last chapter, will again become highly important in expensive. The aim of the smart power grid is to interconnect the finding steady-state voltages and currents for linear circuits having nation within this decade in an efficient power distribution system power sources described by cosine waveforms. A compact nota- that includes conventional and large-scale renewable power tion called phasors will be introduced, and you will perform calcu- sources. Real-time pricing is expected to become commonplace. lations involving complex numbers. Remarkably, dc circuit analysis Smart power grids allocate available power to meet higher pri- techniques that you studied earlier can now be applied directly orities. Wireless sensors, software, and microprocessors are key using phasors for ac circuit analysis. Ac circuits can be analyzed components in this intelligent digital network. As demand for elec- using every basic method—nodal, mesh, superposition, and others. 369 irwin08_369-434hr.qxd 370 28-07-2010 CHAPTER 8 12:03 Page 370 A C S T E A D Y- S TAT E A N A LY S I S 8.1 Let us begin our discussion of sinusoidal functions by considering the sine wave x(t) = XM sin t Sinusoids 8.1 where x(t) could represent either v(t) or i(t). XM is the amplitude, maximum value, or peak value; is the radian or angular frequency; and t is the argument of the sine function. A plot of the function in Eq. (8.1) as a function of its argument is shown in Fig. 8.1a. Obviously, the function repeats itself every 2 radians. This condition is described mathematically as x(t + 2) = x(t), or in general for period T, as x C(t + T)D = x(t) 8.2 meaning that the function has the same value at time t+T as it does at time t. Figure 8.1 x(t) Plots of a sine wave as a function of both t and t. x(t) XM XM — 2 T 3 —— 2 — 2 2 t 3T —— 4 T —— 4 –XM T t –XM (a) (b) The waveform can also be plotted as a function of time, as shown in Fig. 8.1b. Note that this function goes through one period every T seconds. In other words, in 1 second it goes through 1/T periods or cycles. The number of cycles per second, called Hertz, is the frequency f, where [hint] f = The relationship between frequency and period [hint] 1 T 8.3 Now since T = 2, as shown in Fig. 8.1a, we find that = The relationship between frequency, period, and radian frequency 2 = 2f T 8.4 which is, of course, the general relationship among period in seconds, frequency in Hertz, and radian frequency. Now that we have discussed some of the basic properties of a sine wave, let us consider the following general expression for a sinusoidal function: x(t) = XM sin (t + ) [hint] Phase lag defined In this case (t + ) is the argument of the sine function, and is called the phase angle. A plot of this function is shown in Fig. 8.2, together with the original function in Eq. (8.1) for comparison. Because of the presence of the phase angle, any point on the waveform XM sin (t + ) occurs radians earlier in time than the corresponding point on the waveform XM sin t. Therefore, we say that XM sin t lags XM sin (t + ) by radians. In the more general situation, if [hint] In phase and out of phase defined 8.5 x1(t) = XM1 sin (t + ) and x2(t) = XM2 sin (t + ) irwin08_369-434hr.qxd 28-07-2010 12:03 Page 371 SECTION 8.1 SINUSOIDS 371 Figure 8.2 x(t) XM sin (t+) XM sin t 2 Graphical illustration of XM sin˚(t + ) leading XM sin t by radians. t then x1(t) leads x2(t) by - radians and x2(t) lags x1(t) by - radians. If =, the waveforms are identical and the functions are said to be in phase. If Z , the functions are out of phase. The phase angle is normally expressed in degrees rather than radians. Therefore, at this point we will simply state that we will use the two forms interchangeably; that is, x(t) = XM sin a t + b = XM sin (t + 90°) 2 b 2 sin t = cos a t - b 2 8.7 8.8 We are often interested in the phase difference between two sinusoidal functions. Three conditions must be satisfied before we can determine the phase difference: (1) the frequency of both sinusoids must be the same, (2) the amplitude of both sinusoids must be positive, and (3) both sinusoids must be written as sine waves or cosine waves. Once in this format, the phase angle between the functions can be computed as outlined previously. Two other trigonometric identities that normally prove useful in phase angle determination are -cos (t) = cos (t ; 180°) 8.9 -sin (t) = sin (t ; 180°) 8.10 Finally, the angle-sum and angle-difference relationships for sines and cosines may be useful in the manipulation of sinusoidal functions. These relations are sin ( + ) = sin cos + cos sin cos ( + ) = cos cos - sin sin sin ( - ) = sin cos - cos sin cos ( - ) = cos cos + sin sin Phase lead graphically illustrated 8.6 Rigorously speaking, since t is in radians, the phase angle should be as well. However, it is common practice and convenient to use degrees for phase; therefore, that will be our practice in this text. In addition, it should be noted that adding to the argument integer multiples of either 2 radians or 360° does not change the original function. This can easily be shown mathematically but is visibly evident when examining the waveform, as shown in Fig. 8.2. Although our discussion has centered on the sine function, we could just as easily have used the cosine function, since the two waveforms differ only by a phase angle; that is, cos t = sin a t + [hint] 8.11 [hint] A very important point [hint] Some trigonometric identities that are useful in phase angle calculations irwin08_369-434hr.qxd 372 28-07-2010 CHAPTER 8 12:03 Page 372 A C S T E A D Y- S TAT E A N A LY S I S We wish to plot the waveforms for the following functions: EXAMPLE 8.1 SOLUTION a. v(t) = 1 cos (t + 45°), b. v(t) = 1 cos (t + 225°),and c. v(t) = 1 cos (t - 315°). Figure 8.3a shows a plot of the function v(t) = 1 cos t. Figure 8.3b is a plot of the function v(t) = 1 cos (t + 45°).Figure 8.3c is a plot of the function v(t) = 1 cos (t + 225°).Note that since v(t) = 1 cos (t + 225°) = 1 cos (t + 45° + 180°) this waveform is 180° out of phase with the waveform in Fig. 8.3b; that is, cos (t + 225°) = -cos (t + 45°),and Fig. 8.3c is the negative of Fig. 8.3b. Finally, since the function v(t) = 1 cos (t - 315°) = 1 cos (t - 315° + 360°) = 1 cos (t + 45°) this function is identical to that shown in Fig. 8.3b. v(t) v(t) v(t) 1 1 1 45° 135° 90° 180° 270° 360° (a) t –45° 225° 315° t –45° 45° 135° 225° 315° t (c) (b) Figure 8.3 Cosine waveforms with various phase angles. EXAMPLE 8.2 SOLUTION Determine the frequency and the phase angle between the two voltages v1(t) = 12 sin (1000 t + 60°) V and v2(t) = -6 cos (1000t + 30°) V. The frequency in Hertz (Hz) is given by the expression f = 1000 = = 159.2 Hz 2 2 Using Eq. (8.9), v2(t) can be written as v2(t) = -6 cos (t + 30°) = 6 cos (t + 210°) V Then employing Eq. (8.7), we obtain 6 sin (t + 300°) V = 6 sin (t - 60°) V Now that both voltages of the same frequency are expressed as sine waves with positive amplitudes, the phase angle between v1(t) and v2(t) is 60°-(–60°)=120°; that is, v1(t) leads v2(t) by 120° or v2(t) lags v1(t) by 120°. irwin08_369-434hr.qxd 28-07-2010 12:03 Page 373 SECTION 8.2 373 SINUSOIDAL AND COMPLEX FORCING FUNCTIONS Learning Assessments E8.1 Given the voltage v(t) = 120 cos (314t + 兾4) V, determine the frequency of the voltage in Hertz and the phase angle in degrees. E8.2 Three branch currents in a network are known to be ANSWER: f=50 Hz; =45°. ANSWER: i1 leads i2 by –55°; i1 leads i3 by 165°. i1(t) = 2 sin (377t + 45°) A i2(t) = 0.5 cos (377t + 10°) A i3(t) = -0.25 sin (377t + 60°) A Determine the phase angles by which i1(t) leads i2(t) and i1(t) leads i3(t). In the preceding chapters we applied a constant forcing function to a network and found that the steady-state response was also constant. In a similar manner, if we apply a sinusoidal forcing function to a linear network, the steady-state voltages and currents in the network will also be sinusoidal. This should also be clear from the KVL and KCL equations. For example, if one branch voltage is a sinusoid of some frequency, the other branch voltages must be sinusoids of the same frequency if KVL is to apply around any closed path. This means, of course, that the forced solutions of the differential equations that describe a network with a sinusoidal forcing function are sinusoidal functions of time. For example, if we assume that our input function is a voltage v(t) and our output response is a current i(t), as shown in Fig. 8.4, then if v(t) = A sin˚(t + ), i(t) will be of the form i(t) = B sin˚(t + ). The critical point here is that we know the form of the output response, and therefore the solution involves simply determining the values of the two parameters B and . v(t) ± – 8.2 Sinusoidal and Complex Forcing Functions Figure 8.4 Linear electrical network Current response to an applied voltage in an electrical network. i(t) EXAMPLE Consider the circuit in Fig. 8.5. Let us derive the expression for the current. 8.3 R Figure 8.5 v(t)=VM cos t ± – L A simple RL circuit. i(t) SOLUTION The KVL equation for this circuit is L di(t) dt + Ri(t) = VM cos t irwin08_369-434hr.qxd 374 28-07-2010 CHAPTER 8 12:03 Page 374 A C S T E A D Y- S TAT E A N A LY S I S Since the input forcing function is VM cos t, we assume that the forced response component of the current i(t) is of the form i(t) = A cos (t + ) which can be written using Eq. (8.11) as i(t) = A cos cos t - A sin sin t = A 1 cos t + A 2 sin t Note that this is, as we observed in Chapter 7, of the form of the forcing function cos t and its derivative sin t. Substituting this form for i(t) into the preceding differential equation yields L d AA cos t + A 2 sin tB + RAA 1 cos t + A 2 sin tB = VM cos t dt 1 Evaluating the indicated derivative produces -A 1 L sin t + A 2 L cos t + RA 1 cos t + RA 2 sin t = VM cos t By equating coefficients of the sine and cosine functions, we obtain -A 1 L + A 2 R = 0 A 1 R + A 2 L = VM that is, two simultaneous equations in the unknowns A 1 and A 2. Solving these two equations for A 1 and A 2 yields RVM A1 = R + 2L2 LVM 2 A2 = R2 + 2L2 Therefore, i(t) = RVM R2 + 2L2 cos t + LVM R2 + 2L2 which, using the last identity in Eq. (8.11), can be written as i(t) = A cos (t + ) where A and are determined as follows: A cos = A sin = RVM R + 2L2 -LVM 2 R2 + 2L2 Hence, tan = A sin A cos = - L R sin t irwin08_369-434hr.qxd 28-07-2010 12:03 Page 375 SECTION 8.2 SINUSOIDAL AND COMPLEX FORCING FUNCTIONS and therefore, = -tan-1 L R and since (A cos )2 + (A sin )2 = A2 Acos2 + sin2 B = A2 A2 = = A = (L)2 V2M R2V2M AR2 + 2L2 B 2 + AR2 + 2L2 B 2 V2M R2 + 2L2 VM 2R2 + 2L2 Hence, the final expression for i(t) is i(t) = VM 2R + L 2 2 2 cos a t - tan-1 L b R The preceding analysis indicates that is zero if L=0 and hence i(t) is in phase with v(t). If R=0, =–90°, and the current lags the voltage by 90°. If L and R are both present, the current lags the voltage by some angle between 0° and 90°. This example illustrates an important point: solving even a simple one-loop circuit containing one resistor and one inductor is very complicated compared to the solution of a singleloop circuit containing only two resistors. Imagine for a moment how laborious it would be to solve a more complicated circuit using the procedure employed in Example 8.3. To circumvent this approach, we will establish a correspondence between sinusoidal time functions and complex numbers. We will then show that this relationship leads to a set of algebraic equations for currents and voltages in a network (e.g., loop currents or node voltages) in which the coefficients of the variables are complex numbers. Hence, once again we will find that determining the currents or voltages in a circuit can be accomplished by solving a set of algebraic equations; however, in this case, their solution is complicated by the fact that variables in the equations have complex, rather than real, coefficients. The vehicle we will employ to establish a relationship between time-varying sinusoidal functions and complex numbers is Euler’s equation, which for our purposes is written as ejt = cos t + j sin t 8.12 This complex function has a real part and an imaginary part: ReAejt B = cos t ImAejt B = sin t 8.13 where Re(ⴢ) and Im(ⴢ) represent the real part and the imaginary part, respectively, of the function in the parentheses. Recall that j = 1-1. Now suppose that we select as our forcing function in Fig. 8.4 the nonrealizable voltage v(t) = VM ejt 8.14 which because of Euler’s identity can be written as v(t) = VM cos t + jVM sin t 8.15 375 irwin08_369-434hr.qxd 376 28-07-2010 CHAPTER 8 12:03 Page 376 A C S T E A D Y- S TAT E A N A LY S I S The real and imaginary parts of this function are each realizable. We think of this complex forcing function as two forcing functions, a real one and an imaginary one, and as a consequence of linearity, the principle of superposition applies and thus the current response can be written as i(t) = IM cos (t + ) + jIM sin (t + ) 8.16 where IM cos (t + ) is the response due to VM cos t and jIM sin (t + ) is the response due to jVM sin t. This expression for the current containing both a real and an imaginary term can be written via Euler’s equation as i(t) = IM ej(t + ) 8.17 Because of the preceding relationships, we find that rather than apply the forcing function VM cos t and calculate the response IM cos (t + ), we can apply the complex forcing function VM ejt and calculate the response IM ej(t + ), the real part of which is the desired response IM cos (t + ). Although this procedure may initially appear to be more complicated, it is not. It is through this technique that we will convert the differential equation to an algebraic equation that is much easier to solve. EXAMPLE 8.4 SOLUTION Once again, let us determine the current in the RL circuit examined in Example 8.3. However, rather than apply VM cos t, we will apply VM ejt. The forced response will be of the form i(t) = IM ej(t + ) where only IM and are unknown. Substituting v(t) and i(t) into the differential equation for the circuit, we obtain RIM ej(t + ) + L d AI ej(t + ) B = VM ejt dt M Taking the indicated derivative, we obtain RIM ej(t + ) + jLIM ej(t + ) = VM ejt Dividing each term of the equation by the common factor ejt yields RIM ej + jLIM ej = VM which is an algebraic equation with complex coefficients. This equation can be written as IM ej = [hint] Converting the right-hand side of the equation to exponential or polar form produces the equation Summary of complex number relationships: IM ej = x + jy = rej r = 2x2 + y2 y = tan - 1 x x = r cos y = r sin 1 ej = e-j VM R + jL VM -1 2R + L 2 2 2 ejC-tan (L兾R)D (A quick refresher on complex numbers is given in the Appendix for readers who need to sharpen their skills in this area.) The preceding form clearly indicates that the magnitude and phase of the resulting current are IM = VM 2R + 2L2 2 and = -tan -1 L R irwin08_369-434hr.qxd 28-07-2010 12:03 Page 377 SECTION 8.3 PHASORS 377 However, since our actual forcing function was VM cos t rather than VM ejt, our actual response is the real part of the complex response: i(t) = IM cos (t + ) = VM 2R + L 2 2 2 cos a t - tan-1 L b R Note that this is identical to the response obtained in the previous example by solving the differential equation for the current i(t). 8.3 Once again let us assume that the forcing function for a linear network is of the form v(t) = VM ejt Phasors Then every steady-state voltage or current in the network will have the same form and the same frequency ; for example, a current i(t) will be of the form i(t) = IM ej(t + ). As we proceed in our subsequent circuit analyses, we will simply note the frequency and then drop the factor ejt since it is common to every term in the describing equations. Dropping the term ejt indicates that every voltage or current can be fully described by a magnitude and phase. For example, a voltage v(t) can be written in exponential form as v(t) = VM cos (t + ) = ReC VM ej(t + ) D 8.18 v(t) = ReAVM / ejt B 8.19 or as a complex number Since we are working with a complex forcing function, the real part of which is the desired answer, and each term in the equation will contain ejt, we can drop Re(ⴢ) and ejt and work only with the complex number VM / . This complex representation is commonly called a phasor. As a distinguishing feature, phasors will be written in boldface type. In a completely identical manner a voltage v(t) = VM cos (t + ) = ReCVM ej(t + ) D and a current i(t) = IM cos (t + ) = ReCIM ej(t + ) D are written in phasor notation as V = VM / and I = IM / , respectively. Note that it is common practice to express phasors with positive magnitudes. Again, we consider the RL circuit in Example 8.3. Let us use phasors to determine the expression for the current. [hint] If v(t) = VM cos (t + ) and i(t) = IM cos (t + ), then in phasor notation V = VM / and I = IM / EXAMPLE SOLUTION The differential equation is L di(t) dt + Ri(t) = VM cos t The forcing function can be replaced by a complex forcing function that is written as Vejt with phasor V = VM / 0°. Similarly, the forced response component of the current i(t) can be replaced by a complex function that is written as Iejt with phasor I = IM / . From our previous discussions we recall that the solution of the differential equation is the real part of this current. 8.5 irwin08_369-434hr.qxd 378 28-07-2010 CHAPTER 8 12:03 Page 378 A C S T E A D Y- S TAT E A N A LY S I S Using the complex forcing function, we find that the differential equation becomes L d AIejt B + RIejt = Vejt dt jLIejt + RIejt = Vejt Note that ejt is a common factor and, as we have already indicated, can be eliminated, leaving the phasors; that is, jLI + RI = V Therefore, I = [hint] The differential equation is reduced to a phasor equation. VM V L = IM / = -tan-1 2 2 2 ^ R + jL R 2R + L Thus, i(t) = VM 2R + L 2 2 2 cos a t - tan-1 L b R which once again is the function we obtained earlier. We define relations between phasors after the ejt term has been eliminated as “phasor, or frequency domain, analysis.” Thus, we have transformed a set of differential equations with sinusoidal forcing functions in the time domain into a set of algebraic equations containing complex numbers in the frequency domain. In effect, we are now faced with solving a set of algebraic equations for the unknown phasors. The phasors are then simply transformed back to the time domain to yield the solution of the original set of differential equations. In addition, we note that the solution of sinusoidal steady-state circuits would be relatively simple if we could write the phasor equation directly from the circuit description. In Section 8.4 we will lay the groundwork for doing just that. Note that in our discussions we have tacitly assumed that sinusoidal functions would be represented as phasors with a phase angle based on a cosine function. Therefore, if sine functions are used, we will simply employ the relationship in Eq. (8.7) to obtain the proper phase angle. In summary, while v(t) represents a voltage in the time domain, the phasor V represents the voltage in the frequency domain. The phasor contains only magnitude and phase information, and the frequency is implicit in this representation. The transformation from the time domain to the frequency domain, as well as the reverse transformation, is shown in Table 8.1. Recall that the phase angle is based on a cosine function and, therefore, if a sine function is involved, a 90° shift factor must be employed, as shown in the table. TABLE 8.1 Phasor representation TIME DOMAIN FREQUENCY DOMAIN A cos (t ; ) A / ; A sin (t ; ) A / ; - 90° Problem-Solving Strategy Phasor Analysis Step 1. Using phasors, transform a set of differential equations in the time domain into a set of algebraic equations in the frequency domain. Step 2. Solve the algebraic equations for the unknown phasors. Step 3. Transform the now-known phasors back to the time domain. irwin08_369-434hr.qxd 28-07-2010 12:03 Page 379 SECTION 8.4 P H A S O R R E L AT I O N S H I P S F O R C I R C U I T E L E M E N T S 379 However, if a network contains only sine sources, there is no need to perform the 90° shift. We simply perform the normal phasor analysis, and then the imaginary part of the timevarying complex solution is the desired response. Simply put, cosine sources generate a cosine response, and sine sources generate a sine response. Learning Assessments E8.3 Convert the following voltage functions to phasors. ANSWER: V1 = 12 / -425° V; V2 = 18 / -85.8° V. v1(t) = 12 cos (377t - 425°) V v2(t) = 18 sin (2513t + 4.2°) V E8.4 Convert the following phasors to the time domain if the frequency is 400 Hz. ANSWER: v1(t) = 10 cos (800t + 20°) V; v2(t) = 12 cos (800t - 60°) V. V1 = 10 / 20° V V2 = 12 / -60° V As we proceed in our development of the techniques required to analyze circuits in the sinusoidal steady state, we are now in a position to establish the phasor relationships between voltage and current for the three passive elements R, L, and C. In the case of a resistor as shown in Fig. 8.6a, the voltage–current relationship is known to be v(t) = Ri(t) 8.20 Applying the complex voltage VM ejAt + vB results in the complex current IM ejAt + iB, and therefore Eq. (8.20) becomes VM ejAt + vB = RIM ejAt + iB VM ejv = RIM eji 8.21 Equation (8.21) can be written in phasor form as 8.22 where V = VM ejv = VM / v Phasor Relationships for Circuit Elements [hint] which reduces to V = RI 8.4 and I = IM eji = IM / i From Eq. (8.21) we see that v = i and thus the current and voltage for this circuit are in phase. Historically, complex numbers have been represented as points on a graph in which the x-axis represents the real axis and the y-axis the imaginary axis. The line segment connecting the origin with the point provides a convenient representation of the magnitude and angle when the complex number is written in a polar form. A review of the Appendix will indicate how these complex numbers or line segments can be added, subtracted, and so on. Since phasors are complex numbers, it is convenient to represent the phasor voltage and current graphically as line segments. A plot of the line segments representing the phasors is called a phasor diagram. This pictorial representation of phasors provides immediate information on the relative magnitude of one phasor with another, the angle between two phasors, and the relative position of one phasor with respect to another (i.e., leading or lagging). A phasor diagram and the sinusoidal waveforms for the resistor are shown in Figs. 8.6c and d, respectively. A phasor diagram will be drawn for each of the other circuit elements in the remainder of this section. Current and voltage are in phase. irwin08_369-434hr.qxd 380 28-07-2010 CHAPTER 8 12:03 Page 380 A C S T E A D Y- S TAT E A N A LY S I S Figure 8.6 i(t) Voltage–current relationships for a resistor. I + + v(t)=i(t) R R V=RI - - (a) (b) v, i V Im v i I v=i t v=i Re (c) EXAMPLE 8.6 SOLUTION R (d) If the voltage v(t) = 24 cos (377t + 75°) Vis applied to a 6- resistor as shown in Fig. 8.6a, we wish to determine the resultant current. Since the phasor voltage is V = 24 / 75° V the phasor current from Eq. (8.22) is I = 24 / 75° 6 = 4 / 75° A which in the time domain is i(t) = 4 cos (377t + 75°) A Learning Assessment E8.5 The current in a 4- resistor is known to be I = 12 / 60° A. Express the voltage across the resistor as a time function if the frequency of the current is 4 kHz. ANSWER: v(t) = 48 cos (8000t + 60°) V. The voltage–current relationship for an inductor, as shown in Fig. 8.7a, is v(t) = L di(t) dt 8.23 Substituting the complex voltage and current into this equation yields VM ejAt + vB = L d I ejAt + iB dt M which reduces to VM ejv = jLIM eji 8.24 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 381 SECTION 8.4 P H A S O R R E L AT I O N S H I P S F O R C I R C U I T E L E M E N T S i(t) Voltage–current relationships for an inductor. + di(t) v(t)=L ––– dt - V=jLI L L - (a) (b) v(t) v(t), i(t) Im V Figure 8.7 I + 381 i(t) v=i +90° I 90° t i Re (c) (d) Equation (8.24) in phasor notation is [hint] V = jLI The derivative process yields a frequency-dependent function. 8.25 Note that the differential equation in the time domain (8.23) has been converted to an algebraic equation with complex coefficients in the frequency domain. This relationship is shown in Fig. 8.7b. Since the imaginary operator j = 1e j90° = 1 / 90° = 1-1, Eq. (8.24) can be written as VM e jv = LIM e jAi + 90°B 8.26 [hint] The voltage leads the current or the current lags the voltage. Therefore, the voltage and current are 90° out of phase, and in particular the voltage leads the current by 90° or the current lags the voltage by 90°. The phasor diagram and the sinusoidal waveforms for the inductor circuit are shown in Figs. 8.7c and d, respectively. The voltage v(t) = 12 cos (377t + 20°) Vis applied to a 20-mH inductor as shown in Fig. 8.7a. Find the resultant current. EXAMPLE 8.7 SOLUTION The phasor current is I = 12 / 20° V = jL L / 90° [hint] 12 / 20° = (377)(20 * 10 ) / 90° -3 = 1.59 / -70° A Applying V = jLI x1 / 1 x2 / 2 or i(t) = 1.59 cos (377t - 70°) A = x1 / - 2 x2 1 irwin08_369-434hr.qxd 382 28-07-2010 CHAPTER 8 12:03 Page 382 A C S T E A D Y- S TAT E A N A LY S I S Learning Assessment E8.6 The current in a 0.05-H inductor is I = 4 / -30° A. If the frequency of the cur- ANSWER: vL(t) = 75.4 cos (377t + 60°) V. rent is 60 Hz, determine the voltage across the inductor. The voltage–current relationship for our last passive element, the capacitor, as shown in Fig. 8.8a, is i(t) = C dv(t) 8.27 dt Once again employing the complex voltage and current, we obtain IM ejAt + iB = C [hint] d VM ejAt + vB dt which reduces to The current leads the voltage or the voltage lags the current. IM eji = jCVM ejv 8.28 In phasor notation this equation becomes I = jCV 8.29 Eq. (8.27), a differential equation in the time domain, has been transformed into Eq. (8.29), an algebraic equation with complex coefficients in the frequency domain. The phasor relationship is shown in Fig. 8.8b. Substituting j = 1ej90° into Eq. (8.28) yields IM e ji = CVM e jAv + 90°B 8.30 Note that the voltage and current are 90° out of phase. Eq. (8.30) states that the current leads the voltage by 90° or the voltage lags the current by 90°. The phasor diagram and the sinusoidal waveforms for the capacitor circuit are shown in Figs. 8.8c and d, respectively. Figure 8.8 dv(t) i(t)=C ––– dt Voltage–current relationships for a capacitor. I=jCV + + v(t) V C - (a) (b) v(t), i(t) Im I C - i(t) v(t) i=v +90° V 90° t v Re (c) (d) irwin08_369-434hr.qxd 28-07-2010 12:03 Page 383 SECTION 8.5 I M P E D A N C E A N D A D M I T TA N C E 383 EXAMPLE The voltage v(t) = 100 cos (314t + 15°) Vis applied to a 100- F capacitor as shown in Fig. 8.8a. Find the current. 8.8 SOLUTION The resultant phasor current is I = jCA100 / 15° B = (314)A100 * 10-6 / 90°BA100 / 15°B = 3.14 / 105° A [hint] Therefore, the current written as a time function is Applying I = jCV i(t) = 3.14 cos (314t + 105°) A Learning Assessment E8.7 The current in a 150- F capacitor is I = 3.6 / -145° A. If the frequency of the ANSWER: vC(t) = 63.66 cos (377t - 235°) V. current is 60 Hz, determine the voltage across the capacitor. We have examined each of the circuit elements in the frequency domain on an individual basis. We now wish to treat these passive circuit elements in a more general fashion. We define the two-terminal input impedance Z, also referred to as the driving point impedance, in exactly the same manner in which we defined resistance earlier. Later we will examine another type of impedance, called transfer impedance. Impedance is defined as the ratio of the phasor voltage V to the phasor current I: Z = V I 8.5 Impedance and Admittance 8.31 at the two terminals of the element related to one another by the passive sign convention, as illustrated in Fig. 8.9. Since V and I are complex, the impedance Z is complex and Z = VM / v IM / i = VM / v - i = Z / z IM 8.32 Since Z is the ratio of V to I, the units of Z are ohms. Thus, impedance in an ac circuit is analogous to resistance in a dc circuit. In rectangular form, impedance is expressed as Z() = R() + jX() 8.33 where R() is the real, or resistive, component and X() is the imaginary, or reactive, component. In general, we simply refer to R as the resistance and X as the reactance. It is IM VM v ± – Z z Figure 8.9 i ac circuit General impedance relationship. irwin08_369-434hr.qxd 384 28-07-2010 CHAPTER 8 12:03 Page 384 A C S T E A D Y- S TAT E A N A LY S I S important to note that R and X are real functions of and therefore Z() is frequency dependent. Equation (8.33) clearly indicates that Z is a complex number; however, it is not a phasor, since phasors denote sinusoidal functions. Equations (8.32) and (8.33) indicate that Z / z = R + jX 8.34 Therefore, Z = 2R2 + X2 z = tan-1 X R 8.35 where R = Z cos z X = Z sin z For the individual passive elements the impedance is as shown in Table 8.2. However, just as it was advantageous to know how to determine the equivalent resistance in dc circuits, we want to learn how to determine the equivalent impedance in ac circuits. TABLE 8.2 Passive element impedance PASSIVE ELEMENT IMPEDANCE R Z = R L Z = jL = jXL = L / 90° , XL = L C Z = 1 1 1 = jXC = / 90° , XC = jC C C KCL and KVL are both valid in the frequency domain. We can use this fact, as was done in Chapter 2 for resistors, to show that impedances can be combined using the same rules that we established for resistor combinations. That is, if Z 1 , Z 2 , Z 3 , p , Z n are connected in series, the equivalent impedance Z s is Zs = Z1 + Z2 + Z3 + p + Zn 8.36 and if Z 1 , Z 2 , Z 3 , p , Z n are connected in parallel, the equivalent impedance is given by 1 1 1 1 1 = + + + p + Zp Z1 Z2 Z3 Zn EXAMPLE 8.9 8.37 Determine the equivalent impedance of the network shown in Fig. 8.10 if the frequency is f=60 Hz. Then compute the current i(t) if the voltage source is v(t) = 50 cos (t + 30°) V. Finally, calculate the equivalent impedance if the frequency is f=400 Hz. irwin08_369-434hr.qxd 28-07-2010 12:03 Page 385 SECTION 8.5 I M P E D A N C E A N D A D M I T TA N C E 385 Figure 8.10 i(t) Series ac circuit. R=25 v(t) ± – L=20 mH C=50 F SOLUTION The impedances of the individual elements at 60 Hz are Z R = 25 Z L = jL = j(2 * 60)A20 * 10-3 B = j7.54 ZC = -j -j = -j53.05 = C (2 * 60)A50 * 10-6 B Since the elements are in series, Z = ZR + ZL + ZC = 25 - j45.51 The current in the circuit is given by I = 50 / 30° 50 / 30° V = 0.96 / 91.22° A = = Z 25 - j45.51 51.93 / -61.22° or in the time domain, i(t) = 0.96 cos (377t + 91.22°) A. If the frequency is 400 Hz, the impedance of each element is Z R = 25 Z L = jL = j50.27 ZC = -j = -j7.96 C The total impedance is then Z = 25 + j42.31 = 49.14 / 59.42° At the frequency f=60 Hz, the reactance of the circuit is capacitive; that is, if the impedance is written as R+jX, X<0. However, at f=400 Hz the reactance is inductive since X>0. Problem-Solving Strategy Step 1. Express v(t) as a phasor and determine the impedance of each passive element. Step 2. Combine impedances and solve for the phasor I. Step 3. Convert the phasor I to i(t). Basic AC Analysis irwin08_369-434hr.qxd 386 28-07-2010 CHAPTER 8 12:03 Page 386 A C S T E A D Y- S TAT E A N A LY S I S Learning Assessment E8.8 Find the current i(t) in the network in Fig. E8.8. ANSWER: i(t) = 3.88 cos (377t - 39.2°) A. i(t) 20 v(t)=120 sin ± (377t+60°) V – 50 F 40 mH Figure E8.8 Another quantity that is very useful in the analysis of ac circuits is the two-terminal input admittance, which is the reciprocal of impedance; that is, 1 I = Z V Y = 8.38 The units of Y are siemens, and this quantity is analogous to conductance in resistive dc circuits. Since Z is a complex number, Y is also a complex number. Y = YM / y 8.39 which is written in rectangular form as Y = G + jB 8.40 where G and B are called conductance and susceptance, respectively. Because of the relationship between Y and Z, we can express the components of one quantity as a function of the components of the other. From the expression G + jB = [hint] Technique for taking the reciprocal: R - jX 1 = R + jX (R + jX)(R - jX) R - jX = 2 R + X2 1 R + jX 8.41 we can show that G = R 2 R + X 2 , B = , X = -X 2 R + X2 8.42 and in a similar manner, we can show that R = G G 2 + B2 -B G 2 + B2 8.43 It is very important to note that in general R and G are not reciprocals of one another. The same is true for X and B. The purely resistive case is an exception. In the purely reactive case, the quantities are negative reciprocals of one another. The admittance of the individual passive elements are 1 = G R 1 1 YL = = / 90° jL L YR = YC = jC = C / 90° 8.44 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 387 SECTION 8.5 I M P E D A N C E A N D A D M I T TA N C E 387 Once again, since KCL and KVL are valid in the frequency domain, we can show, using the same approach outlined in Chapter 2 for conductance in resistive circuits, that the rules for combining admittances are the same as those for combining conductances; that is, if Y1 , Y2 , Y3 , p , Ynare connected in parallel, the equivalent admittance is Yp = Y1 + Y2 + p + Yn 8.45 and if Y1 , Y2 , p , Ynare connected in series, the equivalent admittance is 1 1 1 1 = + + p + YS Y1 Y2 Yn 8.46 Calculate the equivalent admittance Yp for the network in Fig. 8.11 and use it to determine the current I if VS = 60 / 45° V. I EXAMPLE 8.10 Figure 8.11 VS ± – Yp ZR=2 Example parallel circuit. ZL=j4 SOLUTION From Fig. 8.11 we note that 1 1 = S ZR 2 -j 1 = S YL = ZL 4 YR = [hint] Therefore, 1 1 Yp = - j S 2 4 Admittances add in parallel. and hence, I = Yp VS = a 1 1 - j b A60 / 45°B 2 4 = 33.5 / 18.43° A Learning Assessments E8.9 Find the current I in the network in Fig. E8.9. ANSWER: I = 9.01/53.7° A. I V=10 Figure E8.9 20° V ± – ZR=2 ZL=j2 ZC=–j1 ZR=4 irwin08_369-434hr.qxd 388 28-07-2010 CHAPTER 12:03 8 Page 388 A C S T E A D Y- S TAT E A N A LY S I S As a prelude to our analysis of more general ac circuits, let us examine the techniques for computing the impedance or admittance of circuits in which numerous passive elements are interconnected. The following example illustrates that our technique is analogous to our earlier computations of equivalent resistance. Consider the network shown in Fig. 8.12a. The impedance of each element is given in the figure. We wish to calculate the equivalent impedance of the network Z eq at terminals A–B. EXAMPLE 8.11 1 4 A j2 2 –j2 j6 Zeq j4 –j2 A Z1 Zeq – j2 B Z3 Z2 Z4 B (a) (b) Figure 8.12 Example circuit for determining equivalent impedance in two steps. SOLUTION The equivalent impedance Z eq could be calculated in a variety of ways; we could use only impedances, or only admittances, or a combination of the two. We will use the latter. We begin by noting that the circuit in Fig. 8.12a can be represented by the circuit in Fig. 8.12b. Note that Y4 = YL + YC = 1 1 + j4 -j2 = j 1 S 4 Therefore, Z 4 = -j4 Now Z 34 = Z 3 + Z 4 = (4 + j2) + (-j4) = 4 - j2 and hence, Y34 = = 1 Z 34 1 4 - j2 = 0.20 + j0.10 S irwin08_369-434hr.qxd 28-07-2010 12:03 Page 389 SECTION 8.5 I M P E D A N C E A N D A D M I T TA N C E 389 Since Z 2 = 2 + j6 - j2 = 2 + j4 then 1 2 + j4 Y2 = = 0.10 - j0.20 S Y234 = Y2 + Y34 = 0.30 - j0.10 S The reader should carefully note our approach: we are adding impedances in series and adding admittances in parallel. From Y234 we can compute Z 234 as 1 Y234 Z 234 = 1 0.30 - j0.10 = = 3 + j1 Now Y1 = YR + YC = 1 1 + 1 -j2 = 1 + j 1 S 2 and then Z1 = 1 1 2 = 0.8 - j0.4 1 + j Therefore, Z eq = Z 1 + Z 234 = 0.8 - j0.4 + 3 + j1 = 3.8 + j0.6 Problem-Solving Strategy Step 1. Add the admittances of elements in parallel. Step 2. Add the impedances of elements in series. Step 3. Convert back and forth between admittance and impedance in order to combine neighboring elements. Combining Impedances and Admittances irwin08_369-434hr.qxd 390 28-07-2010 CHAPTER 8 12:03 Page 390 A C S T E A D Y- S TAT E A N A LY S I S Learning Assessments ANSWER: Z T = 3.38 + j1.08 . E8.10 Compute the impedance Z T in the network in Fig. E8.10. 2 –j4 j2 4 2 j6 ZT Figure E8.10 E8.11 Find Z in Fig. E8.11. 2 ANSWER: Z = 1.95 + j0.29 . j3 4 j2 3 5 Z –j3 Figure E8.11 8.6 Phasor Diagrams EXAMPLE 4 –j3 Impedance and admittance are functions of frequency, and therefore their values change as the frequency changes. These changes in Z and Y have a resultant effect on the current–voltage relationships in a network. This impact of changes in frequency on circuit parameters can be easily seen via a phasor diagram. The following examples will serve to illustrate these points. Let us sketch the phasor diagram for the network shown in Fig. 8.13. V 8.12 IS Figure 8.13 IR 1 — jL R IL jC IC Example parallel circuit. SOLUTION The pertinent variables are labeled on the figure. For convenience in forming a phasor diagram, we select V as a reference phasor and arbitrarily assign it a 0° phase angle. We will, therefore, measure all currents with respect to this phasor. We suffer no loss of generality by assigning V a 0° phase angle, since if it is actually 30°, for example, we will simply rotate the entire phasor diagram by 30° because all the currents are measured with respect to this phasor. At the upper node in the circuit KCL is IS = IR + IL + IC = V V V + + R jL 1兾jC irwin08_369-434hr.qxd 28-07-2010 12:03 Page 391 SECTION 8.6 PHASOR DIAGRAMS 391 Figure 8.14 IC Phasor diagrams for the circuit in Fig. 8.13. IC IR IR V V IL+IC IL IS IL (a) (b) I Sn IC ISf IS IC+IL IR IR V V IL IS2 (c) (d) IS1 Since V = VM / 0°, then IS = VM / 0° R VM / -90° + L + VM C / 90° The phasor diagram that illustrates the phase relationship between V, IR , IL and IC is shown in Fig. 8.14a. For small values of such that the magnitude of IL is greater than that of IC, the phasor diagram for the currents is shown in Fig. 8.14b. In the case of large values of —that is, those for which IC is greater than IL—the phasor diagram for the currents is shown in Fig. 8.14c. Note that as increases, the phasor I S moves from I S1 to I Sn along a locus of points specified by the dashed line shown in Fig. 8.14d. Note that I S is in phase with V when I C = I L or, in other words, when L=1/C. Hence, the node voltage V is in phase with the current source I S when = 1 1LC [hint] This can also be seen from the KCL equation I = c 1 1 b dV + j a C R L Let us determine the phasor diagram for the series circuit shown in Fig. 8.15a. KVL for this circuit is of the form EXAMPLE SOLUTION VS = VR + VL + VC = IR + LI / 90° + From a graphical standpoint, phasors can be manipulated like vectors. I / -90° C 8.13 irwin08_369-434hr.qxd 392 28-07-2010 CHAPTER 8 12:03 Page 392 A C S T E A D Y- S TAT E A N A LY S I S If we select I as a reference phasor so that I = IM / 0°, then if LIM 7 IM兾C, the phasor diagram will be of the form shown in Fig. 8.15b. Specifically, if =377 rad/s (i.e., f=60 Hz), then L=6 and 1/C=2. Under these conditions, the phasor diagram is as shown in Fig. 8.15c. If, however, we select VS as reference with, for example, vS(t) = 12 12 cos (377t + 90°) V then I = 1212 / 90° V = Z 4 + j6 - j2 1212 / 90° = 422 / 45° = 3 / 45° A and the entire phasor diagram, as shown in Figs. 8.15b and c, is rotated 45°, as shown in Fig. 8.15d. Figure 8.15 VL Series circuit and certain specific phasor diagrams (plots are not drawn to scale). R=4 I + VR- ± VS – + VL - VC + VS VL+VC L=15.92 mH VR - I VC C=1326 F (a) (b) VL=6IM/90° VL VS I VS VL+VC 45° VL+VC I 45° VR=4IM 0° VC=2IM –90° 45° VR VC (c) (d) Learning Assessments E8.12 Draw a phasor diagram to illustrate all currents and voltages ANSWER: for the network in Fig. E8.12. I=4 A V=7.16 V + I=4 45° A 2 I1 Figure E8.12 –j4 I2 V - I2=1.79 A 108° 45° 18.43° I1=3.58 A irwin08_369-434hr.qxd 28-07-2010 12:03 Page 393 SECTION 8.7 B A S I C A N A LY S I S U S I N G K I R C H H O F F ' S L A W S E8.13 Find the value of C such that v(t) and i(t) are in phase in Fig. E8.13. 393 ANSWER: C = 400 µF + i(t) Figure E8.13 v(t) 4 C 10 mH - We have shown that Kirchhoff’s laws apply in the frequency domain, and therefore they can be used to compute steady-state voltages and currents in ac circuits. This approach involves expressing these voltages and currents as phasors, and once this is done, the ac steady-state analysis employing phasor equations is performed in an identical fashion to that used in the dc analysis of resistive circuits. Complex number algebra is the tool that is used for the mathematical manipulation of the phasor equations, which, of course, have complex coefficients. We will begin by illustrating that the techniques we have applied in the solution of dc resistive circuits are valid in ac circuit analysis also—the only difference being that in steady-state ac circuit analysis the algebraic phasor equations have complex coefficients. 8.7 Basic Analysis Using Kirchhoff’s Laws Problem-Solving Strategy ● For relatively simple circuits (e.g., those with a single source), use ● Ohm’s law for ac analysis—that is, V = IZ ● The rules for combining Z s and Yp KCL and KVL ● Current and voltage division For more complicated circuits with multiple sources, use ● Nodal analysis ● Loop or mesh analysis ● Superposition ● Source exchange ● Thévenin’s and Norton’s theorems ● ● At this point, it is important for the reader to understand that in our manipulation of algebraic phasor equations with complex coefficients we will, for the sake of simplicity, normally carry only two digits to the right of the decimal point. In doing so, we will introduce round-off errors in our calculations. Nowhere are these errors more evident than when two or more approaches are used to solve the same problem, as is done in the following example. AC Steady-State Analysis irwin08_369-434hr.qxd 394 28-07-2010 CHAPTER 8 Page 394 A C S T E A D Y- S TAT E A N A LY S I S We wish to calculate all the voltages and currents in the circuit shown in Fig. 8.16a. EXAMPLE 8.14 12:03 SOLUTION Our approach will be as follows. We will calculate the total impedance seen by the source VS . Then we will use this to determine I 1 . Knowing I 1 , we can compute V1 using KVL. Knowing V1 , we can compute I 2 and I 3 , and so on. The total impedance seen by the source VS is Z eq = 4 + = 4 + (j6)(8 - j4) j6 + 8 - j4 24 + j48 8 + j2 = 4 + 4.24 + j4.94 = 9.61 / 30.94° [hint] I1 4 8 I3 I3 Technique 1. Compute I1 . VS=24 I1 VS 60° V ± – + j6 V1 - 105° + V2 I1 –j4 - I2 29.06° –11.58° ± – I2 (a) I1 VS ± – Figure 8.16 (a) Example ac circuit, (b) phasor diagram for the currents Z1 I2 + V1 (b) - Z2 I3 Z3 (plots are not drawn to scale). Then I1 = 2. Determine V1 = Vs - I1Z1 Then I2 = = 2.5 / 29.06° A V1 V1 and I3 = Z2 Z3 Current and voltage division are also applicable. 24 / 60° VS = Z eq 9.61 / 30.94° V1 can be determined using KVL: V1 = VS - 4I 1 = 24 / 60° - 10 / 29.06° = 3.26 + j15.93 = 16.26 / 78.43° V Note that V1 could also be computed via voltage division: VS V1 = 4 + which from our previous calculation is (j6)(8 - j4) j6 + 8 - j4 (j6)(8 - j4) j6 + 8 - 4 V irwin08_369-434hr.qxd 28-07-2010 12:03 Page 395 SECTION 8.7 V1 = B A S I C A N A LY S I S U S I N G K I R C H H O F F ' S L A W S 395 A24 / 60°B A6.51 / 49.36°B 9.61 / 30.94° = 16.26 / 78.42° V Knowing V1 , we can calculate both I 2 and I 3: I2 = 16.26 / 78.43° V1 = j6 6 / 90° = 2.71 / -11.58° A and I3 = V1 8 - j4 = 1.82 / 105° A Note that I 2 and I 3 could have been calculated by current division. For example, I 2 could be determined by I2 = = I 1(8 - j4) 8 - j4 + j6 A2.5 / 29.06°BA8.94 / -26.57°B 8 + j2 = 2.71 / -11.55° A Finally, V2 can be computed as V2 = I 3(-j4) = 7.28 / 15° V This value could also have been computed by voltage division. The phasor diagram for the currents I 1 , I 2 , and I 3 is shown in Fig. 8.16b and is an illustration of KCL. Finally, the reader is encouraged to work the problem in reverse; that is, given V2 , find VS . Note that if V2 is known, I 3 can be computed immediately using the capacitor impedance. Then V2 + I 3(8) yields V1 . Knowing V1 we can find I 2 . Then I 2 + I 3 = I 1 , and so on. Note that this analysis, which is the subject of Learning Assessment E8.12, involves simply a repeated application of Ohm’s law, KCL, and KVL. Learning Assessments ANSWER: vA(t)= 95.83 cos(50t+24.1°) V. E8.14 Find vA(t) in Fig. E8.14. 2 0.01 F 0.05 H + 60 cos (50t+20°) V + – + – vA(t) 0.1 H Figure E8.14 5 0.005 F 100 cos 50t V irwin08_369-434hr.qxd 396 28-07-2010 CHAPTER 12:03 8 Page 396 A C S T E A D Y- S TAT E A N A LY S I S ANSWER: Vo = 2.98 / -153.43° V. E8.15 Find Vo in Fig. E8.15. 4 2 + –j 2 2 0° A j 2 –j 2 Vo - Figure E8.15 E8.16 In the network in Fig. E8.16, Vo is known to be 8 / 45° V. Compute VS. I1 I3 2 VS ± – ANSWER: VS = 17.89 / -18.43° V. + V1 - I2 j2 –j2 + 2 Vo - Figure E8.16 8.8 Analysis Techniques EXAMPLE 8.15 SOLUTION In this section we revisit the circuit analysis methods that were successfully applied earlier to dc circuits and illustrate their applicability to ac steady-state analysis. The vehicle we employ to present these techniques is examples in which all the theorems, together with nodal analysis and loop analysis, are used to obtain a solution. Let us determine the current I o in the network in Fig. 8.17a using nodal analysis, loop analysis, superposition, source exchange, Thévenin’s theorem, and Norton’s theorem. 1. Nodal Analysis We begin with a nodal analysis of the network. The KCL equation for the supernode that includes the voltage source is [hint] Summing the current, leaving the supernode. Outbound currents have a positive sign. V1 V2 V2 - 2 / 0° + + = 0 1 + j 1 1 - j irwin08_369-434hr.qxd 28-07-2010 12:03 Page 397 SECTION 8.8 j1 V1 6 0° V + 1 – j1 1 1 1 I2 2 0° A (a) 397 1 + I1 Io 2 0° A 6 0° V j1 1 V2 A N A LY S I S T E C H N I Q U E S I3 –j1 Io (b) Figure 8.17 Circuits used in Example 8.15 for node and loop analysis. and the associated KVL constraint equation is V1 + 6 / 0° = V2 Solving for V1 in the second equation and using this value in the first equation yields V2 - 6 / 0° 1 + j - 2 / 0° + V2 + V2 = 0 1 - j or V2 c 6 + 2 + 2j 1 1 + 1 + d = 1 + j 1 - j 1 + j Solving for V2 , we obtain V2 = a 4 + j bV 1 + j Therefore, Io = 4 + j 3 5 = a - jb A 1 + j 2 2 2. Loop Analysis The network in Fig. 8.17b is used to perform a loop analysis. Note that one loop current is selected that passes through the independent current source. The three loop equations are I 1 = -2 / 0° 1AI 1 + I 2 B + j1AI 1 + I 2 B - 6 / 0° + 1AI 2 + I 3 B - j1AI 2 + I 3 B = 0 1I 3 + 1AI 2 + I 3 B - j1AI 2 + I 3 B = 0 Combining the first two equations yields I 2(2) + I 3(1 - j) = 8 + 2j The third loop equation can be simplified to the form I 2(1 - j) + I 3(2 - j) = 0 Solving this last equation for I 2 and substituting the value into the previous equation yields I3 c -4 + 2j + 1 - j d = 8 + 2j 1 - j or I3 = -10 + 6j 4 [hint] Just as in a dc analysis, the loop equations assume that a decrease in potential level is and an increase is . irwin08_369-434hr.qxd 398 28-07-2010 CHAPTER 8 12:03 Page 398 A C S T E A D Y- S TAT E A N A LY S I S and finally I o = -I 3 = a 5 3 - jb A 2 2 3. Superposition In using superposition, we apply one independent source at a time. [hint] In applying superposition in this case, each source is applied independently, and the results are added to obtain the solution. The network in which the current source acts alone is shown in Fig. 8.18a. By combining the two parallel impedances on each end of the network, we obtain the circuit in Fig. 8.18b, where Z¿ = (1 + j)(1 - j) (1 + j) + (1 - j) = 1 Therefore, using current division, I oœ = 1 / 0° A The circuit in which the voltage source acts alone is shown in Fig. 8.18c. The voltage V1fl obtained using voltage division is V1fl (6 / 0°) c 1(1 - j) 1 + 1 - j d = 1(1 - j) 1 + j + c 6(1 - j) = 4 1 + 1 - j d V and hence, I 0fl = 6 (1 - j) A 4 Then I o = I oœ + I ofl = 1 + 3 6 5 (1 - j) = a - j b A 4 2 2 4. Source Exchange As a first step in the source exchange approach, we exchange the current source and parallel impedance for a voltage source in series with the impedance, as shown in Fig. 8.19a. j1 Figure 8.18 Circuits used in Example 8.15 for a superposition analysis. 1 1 1 2 0° A –j1 2 0° A 1 I'o I'o (a) j1 (b) 1 6 0° V + + 1 V"1 - (c) 1 I"o –j1 Z' irwin08_369-434hr.qxd 28-07-2010 12:03 Page 399 SECTION 8.8 + j1 1 A N A LY S I S T E C H N I Q U E S Figure 8.19 1 Circuits used in Example 8.15 for a source exchange analysis. 6 0° V ± – 1 2(1+j) V 399 –j1 Io (a) 1 1 6+2(1+j) c— c A 1+j 1 –j1 j1 Io 8+2j c— c A 1+j (c) Adding the two voltage sources and transforming them and the series impedance into a current source in parallel with that impedance are shown in Fig. 8.19b. Combining the two impedances that are in parallel with the 1- resistor produces the network in Fig. 8.19c, where (1 + j)(1 - j) 1 + j + 1 - j = 1 Therefore, using current division, Io = a = a [hint] In source exchange, a voltage source in series with an impedance can be exchanged for a current source in parallel with the impedance and vice versa. Repeated application systematically reduces the number of circuit elements. 8 + 2j 4 + j 1 ba b = 1 + j 2 1 + j 5 3 - jb A 2 2 5. Thévenin Analysis In applying Thévenin’s theorem to the circuit in Fig. 8.17a, we first find the open-circuit voltage, Voc , as shown in Fig. 8.20a. To simplify the analysis, we perform a source exchange on the left end of the network, which results in the circuit in Fig. 8.20b. Now using voltage division, Voc = [6 + 2(1 + j)] c 1 - j d 1 - j + 1 + j or Voc = (5 - 3j) V The Thévenin equivalent impedance, Z Th , obtained at the open-circuit terminals when the current source is replaced with an open circuit and the voltage source is replaced with a short circuit is shown in Fig. 8.20c and calculated to be Z Th = Z Io (b) Z = 1 (1 + j)(1 - j) 1 + j + 1 - j = 1 [hint] In this Thévenin analysis, 1. Remove the 1- load and find the voltage across the open terminals, Voc. 2. Determine the impedance ZTh at the open terminals with all sources made zero. 3. Construct the following circuit and determine Io. ZTh Voc ± – Io 1 irwin08_369-434hr.qxd 400 28-07-2010 CHAPTER 12:03 8 Page 400 A C S T E A D Y- S TAT E A N A LY S I S 6 0° V 6 0° V + j1 + 1 2 0° A j1 1 ± – – j1 Voc 1 + + 2(1+j) V 1 –j1 Voc - - (a) (b) ZTh=1 j1 1 Voc=(5-3j) V ± – –j1 ZTh 1 1 Io Figure 8.20 (d) (c) Circuits used in Example 8.15 for a Thévenin analysis. [hint] Connecting the Thévenin equivalent circuit to the 1- resistor containing I o in the original network yields the circuit in Fig. 8.20d. The current I o is then In this Norton analysis, 1. Remove the 1- load and find the current Isc through the short-circuited terminals. 2. Determine the impedance ZTh at the open load terminals with all sources made zero. 3. Construct the following circuit and determine Io. Io = a 6. Norton Analysis Finally, in applying Norton’s theorem to the circuit in Fig. 8.17a, we calculate the short-circuit current, I sc , using the network in Fig. 8.21a. Note that because of the short circuit, the voltage source is directly across the impedance in the left-most branch. Therefore, 6 / 0° I1 = ZTh 1 + j Then, using KCL, Io Isc 5 3 - jb A 2 2 I sc = I 1 + 2 / 0° = 2 + 1 = a 6 1 + j 8 + 2j bA 1 + j Figure 8.21 Circuits used in Example 8.15 for a Norton analysis. 6 0° V j1 + 1 Isc 1 2 0° A I1 (a) –j1 8+2j Isc=r—r A 1+j 1 ZTh=1 (b) Io irwin08_369-434hr.qxd 28-07-2010 12:03 Page 401 SECTION 8.8 A N A LY S I S T E C H N I Q U E S 401 The Thévenin equivalent impedance, ZTh, is known to be 1 and, therefore, connecting the Norton equivalent to the 1- resistor containing I o yields the network in Fig. 8.21b. Using current division, we find that Io = 1 8 + 2j a b 2 1 + j = a 5 3 - jb A 2 2 Let us now consider an example containing a dependent source Let us determine the voltage Vo in the circuit in Fig. 8.22a. In this example we will use node equations, loop equations, Thévenin’s theorem, and Norton’s theorem. We will omit the techniques of superposition and source transformation. Why? 1. Nodal Analysis To perform a nodal analysis, we label the node voltages and identify the EXAMPLE 8.16 SOLUTION supernode as shown in Fig. 8.22b. The constraint equation for the supernode is V3 + 12 / 0° = V1 V1 –j 1 12 0° V ± – 4 0° A 1 1 –j 1 Ix j1 1 1 V2 + 2Ix 12 0° V ± – V3 1 Ix + 2Ix Vo 4 0° A j1 - 1 Vo - (a) (b) –j 1 I1 I2 ± – 12 0° V Ix 1 2 Ix 1 I3 j1 (c) 4 0° A I4 1 + Vo - Figure 8.22 Circuits used in Example 8.16 for nodal and loop analysis. irwin08_369-434hr.qxd 402 28-07-2010 CHAPTER [hint] 8 12:03 Page 402 A C S T E A D Y- S TAT E A N A LY S I S and the KCL equations for the nodes of the network are V1 - V2 V3 - V2 V3 V3 - Vo + - 4 / 0° + + = 0 -j1 1 1 j1 How does the presence of a dependent source affect superposition and source exchange? V3 - Vo V2 - V1 V2 - V3 + - 2a b = 0 -j1 1 1 4 / 0° + Vo Vo - V3 + = 0 1 1 At this point we can solve the foregoing equations using a matrix analysis or, for example, substitute the first and last equations into the remaining two equations, which yields 3Vo - (1 + j)V2 = -(4 + j12) -(4 + j2)Vo + (1 + j)V2 = 12 + j16 Solving these equations for Vo yields Vo = -(8 + j4) 1 + j2 = +4 / 143.13° V 2. Loop Analysis The mesh currents for the network are defined in Fig. 8.22c. The constraint equations for the circuit are I 2 = -4 / 0° I x = I 4 - I 2 = I 4 + 4 / 0° I 3 = 2I x = 2I 4 + 8 / 0° The KVL equations for mesh 1 and mesh 4 are -j1I 1 + 1AI 1 - I 3 B = -12 / 0° j1AI 4 - I 3 B + 1AI 4 - I 2 B + 1I 4 = 0 Note that if the constraint equations are substituted into the second KVL equation, the only unknown in the equation is I 4 . This substitution yields I 4 = +4 / 143.13° A and hence, Vo = +4 / 143.13° V 3. Thévenin’s Theorem In applying Thévenin’s theorem, we will find the open-circuit voltage and then determine the Thévenin equivalent impedance using a test source at the open-circuit terminals. We could determine the Thévenin equivalent impedance by calculating the short-circuit current; however, we will determine this current when we apply Norton’s theorem. The open-circuit voltage is determined from the network in Fig. 8.23a. Note that I xœ = 4 / 0° A and since 2I xœ flows through the inductor, the open-circuit voltage Voc is Voc = -1A4 / 0°B + j1A2I xœ B = -4 + j8 V To determine the Thévenin equivalent impedance, we turn off the independent sources, apply a test voltage source to the output terminals, and compute the current leaving the test source. As shown in Fig. 8.23b, since I xfl flows in the test source, KCL requires irwin08_369-434hr.qxd 28-07-2010 12:03 Page 403 SECTION 8.8 –j1 12 0° V ± – 1 403 –j 1 4 0° A 1 A N A LY S I S T E C H N I Q U E S 1 I'x 1 I"x + j1 2Ix 2I"x Voc j1 ± V test – I"x (a) (b) 1 + –j1 –4+j8 V ± – 1 Vo - (c) that the current in the inductor be I xfl also. KVL around the mesh containing the test source indicates that j1I xfl - 1I xfl - Vtest = 0 Therefore, I xfl = -Vtest 1 - j Then Z Th = Vtest -I xfl = 1 - j If the Thévenin equivalent network is now connected to the load, as shown in Fig. 8.23c, the output voltage Vo is found to be Vo = -4 + 8j (1) 2 - j1 = +4 / 143.13° V 4. Norton’s Theorem In using Norton’s theorem, we will find the short-circuit current from the network in Fig. 8.24a. Once again, using the supernode, the constraint and KCL equations are V3 + 12 / 0° = V1 V2 - V1 V2 - V3 + - 2I Ô x = 0 -j1 1 V1 - V2 V3 - V2 V3 + - 4 / 0° + + IÔ x = 0 -j1 1 j1 IÔ x = V3 1 Figure 8.23 Circuits used in Example 8.16 when applying Thévenin’s theorem. irwin08_369-434hr.qxd 404 28-07-2010 CHAPTER 8 12:03 Page 404 A C S T E A D Y- S TAT E A N A LY S I S V1 + –j 1 V2 12 0° V ± – 1 4 0° A I'"x V3 1 –(8+j4) —A 1+j 1 2I'"x j1 1 Isc –j1 Vo Io - (a) (b) Figure 8.24 Circuits used in Example 8.16 when applying Norton’s theorem. Substituting the first and last equations into the remaining equations yields (1 + j)V2 - (3 + j)I Ô x = j12 -(1 + j)V2 + (2)I Ô x = 4 - j12 Solving these equations for I Ô x yields IÔ x = -4 A 1 + j The KCL equation at the right-most node in the network in Fig. 8.24a is IÔ x = 4 / 0° + I sc Solving for I sc , we obtain I sc = -(8 + j4) 1 + j A The Thévenin equivalent impedance was found earlier to be Z Th = 1 - j Using the Norton equivalent network, the original network is reduced to that shown in Fig. 8.24b. The voltage Vo is then Vo = -(8 + j4) (1)(1 - j) c d 1 + j 1 + 1 - j = -4 c 3 - j d 3 + j = +4 / +143.13° V irwin08_369-434hr.qxd 28-07-2010 12:03 Page 405 SECTION 8.8 A N A LY S I S T E C H N I Q U E S 405 Learning Assessments ANSWER: Vo = 2.12 / 75° V. E8.17 Use nodal analysis to find Vo in the network in Fig. E8.17. 2 – j1 + 12 30° V ± – j2 1 1 Vo - Figure E8.17 ANSWER: I 1 = 0.7781 / -161.9° A. E8.18 Find I1 in Fig. E8.18 using nodal analysis. j2 I1 3 4 j3 + – –j2 12 0° V 6 0° V 3 10° A + – Figure E8.18 E8.19 Find Vx in Fig. E8.19 using (a) nodal analysis and (b) mesh analysis. ANSWER: Vx = 17.4 / -21.62° V. 4 + 24 0° V + – 3 j4 j2 Vx 4 10 0° V + – –j2 - Figure E8.19 E8.20 Use (a) mesh equations and (b) Thévenin’s theorem to find Vo in the network in the network in Fig. E8.20. 2 – j2 + j2 24 0° V ± – I1 I2 2 Vo 2 90° A Figure E8.20 - ANSWER: Vo = 10.88 / 36° V. irwin08_369-434hr.qxd 406 28-07-2010 CHAPTER 8 12:03 Page 406 A C S T E A D Y- S TAT E A N A LY S I S ANSWER: Vo = 1.4654 / -12.34° V. E8.21 Find Vo in Fig. E8.21 using mesh analysis. 4 24 0° V + – j4 3 j2 + + 6 4 VA + – – Figure E8.21 2 VA Vo –j2 – E8.22 Find I1 in Fig. E8.18 using superposition ANSWER: I 1 = 0.7781 / -161.9° V. j2 I1 3 4 j3 + – – j2 12 0° V 6 0° V 3 10° A + – Figure E8.22 E8.23 Use (a) superposition, (b) source transformation, and (c) Norton’s theorem to find ANSWER: Vo = 12 / 90° V. Vo in the network in Fig. E8.23. 2 –j2 + j2 24 0° V Figure E8.23 ± – 2 + Vo 12 0° V - irwin08_369-434hr.qxd 28-07-2010 12:03 Page 407 SECTION 8.8 A N A LY S I S T E C H N I Q U E S 407 ANSWER: Vo = 1.4654 / -12.34° V. E8.24 Find Vo in Figure E8.24 using Thevenin’s theorem. 4 24 0° V + – j4 3 j2 + + 6 4 VA + – – Figure E8.24 2 VA Vo –j2 – Let’s solve for the current i(t) in the circuit in Fig. 8.25. At first glance, this appears to be a simple single-loop circuit. A more detailed observation reveals that the two sources operate at different frequencies. The radian frequency for the source on the left is 10 rad兾s, while the source on the right operates at a radian frequency of 20 rad兾s. If we draw a frequency-domain circuit, which frequency do we use? How can we solve this problem? Recall that the principle of superposition tells us that we can analyze the circuit with each source operating alone. The circuit responses to each source acting alone are then added together to give us the response with both sources active. Let’s use the principle of superposition to solve this problem. First, calculate the response i¿(t) from the source on the left using the circuit shown in Fig. 8.26a. Now we can draw a frequency-domain circuit for = 10 rad兾s. 10 100 cos 10t V 100 / 0° ± – 1H 8.17 SOLUTION Figure 8.25 i(t) ± – Circuit used in Example 8.17. 50 cos (20t-10°) V = 7.07 / -45° A. Therefore, i¿(t) = 7.07 cos (10t - 45°) A. 10 + j10 The response due to the source on the right can be determined using the circuit in Fig. 8.27. Note that i–(t) is defined in the opposite direction to i(t) in the original circuit. The frequency-domain circuit for = 20 rad兾s is also shown in Fig. 8.27b. 50 / -10° = 2.24 / -73.43° A. Therefore, i–(t) = 2.24 cos (20t - 73.43°) A. The current I– = 10 + j20 The current i(t) can now be calculated as i¿(t) - i–(t) = 7.07 cos (10t - 45°) 2.24 cos (20t - 73.43°) A. Then I¿ = EXAMPLE irwin08_369-434hr.qxd 408 28-07-2010 CHAPTER 8 12:03 A C S T E A D Y- S TAT E A N A LY S I S 10 100 cos 10t V Page 408 1H i'(t) 10 i"(t) 1H ± – (a) I' j20 10 ± – 50 –10° V I" ± – 100 0° V 50 cos (20t-10°) V (a) j10 10 ± – (b) Figure 8.26 Circuits used to illustrate superposition. (b) Figure 8.27 Circuits used to illustrate superposition. 8.9 Application Examples APPLICATION EXAMPLE 8.18 The network in Fig. 8.28 models an unfortunate situation that is all too common. Node A, which is the voltage vin(t) at the output of a temperature sensor, has “picked-up” a highfrequency voltage, vnoise(t), caused by a nearby AM radio station. The noise frequency is 700 kHz. In this particular scenario, the sensor voltage, like temperature, tends to vary slowly. Our task then is to modify the circuit to reduce the noise at the output without disturbing the desired signal, vin(t). Figure 8.28 Modeling radio frequency noise pickup. A vin(t) ± – vnoise(t) ± – + R 10 k vo(t) - irwin08_369-434hr.qxd 28-07-2010 12:03 Page 409 SECTION 8.9 X + vin(t) ± – vnoise(t) ± – jL A R 10 k Vin ± – Vnoise ± – vo(t) - A P P L I C AT I O N E X A M P L E S Figure 8.29 A + R 10 k 409 Vo (a) Model used to reject Vnoise and (b) the required component. (b) (a) Consider the network in Fig. 8.29a. If component X has a high impedance—that is, much greater than R at 700 kHz and an impedance of zero at dc—we should be able to alleviate the problem. Using voltage division to obtain Vo in Fig. 8.29b, we find Vo = c SOLUTION R d V1 R + jL where V1 and are either Vin and 0, or Vnoise and 2(700 * 103). At dc, = 0, the inductor’s impedance is zero, the voltage division ratio is unity, V1 is Vin and Vo equals Vin . But at 700 kHz, V1 is Vnoise and the desired voltage division ratio should be very small; that is, the inductor impedance must be much greater than R, so that Vo becomes nearly zero. If we choose to reduce the noise at the output by 90%, we find ` R 1 ` = R + jL 10 at f = 700 kHz Solving this equation yields L = 22.6 mH,which is close to a standard inductor value. The circuit in Fig. 8.30 is called a General Impedance Converter (GIC). We wish to develop an expression for the impedance Z eq in terms of Z 1 , Z 2 , Z 3 , Z 4, and Z 5 , and then using resistors of equal value and a 1- F capacitor, create a 1-H equivalent inductance. Vin Figure 8.30 Zeq The general impedance converter. Iin E V3 Vin D Z4 C B Vin ± – Z5 Z3 + APPLICATION EXAMPLE 8.19 Z2 A Z1 V2 irwin08_369-434hr.qxd 410 28-07-2010 CHAPTER 8 SOLUTION 12:03 Page 410 A C S T E A D Y- S TAT E A N A LY S I S We simply employ the ideal op-amp assumptions; that is, there is no current entering any op-amp input, and the voltage across any op-amp’s input terminals is zero. As a result, the voltage at both nodes A and C are Vin . Next, we apply KCL at each op-amp input. At node A V2 - Vin Vin = Z2 Z1 which yields V2 = Vin c 1 + Z2 d Z1 8.47 At node C, we find V3 - Vin Vin - V2 = Z4 Z3 Solving for V3 yields V3 = Vin c 1 + Z4 Z4 d - V2 c d Z3 Z3 Substituting our results from Eq. (8.47), we can express V3 as V3 = Vin c 1 - Z2 Z4 d Z1 Z3 8.48 At node E, we write I in = Vin - V3 Z5 Then using our expression for V3 in Eq. (8.48), we find that I in = Vin c Z2 Z4 d Z1 Z3 Z5 Finally, the impedance of interest is Z eq = Z1 Z3 Z5 Vin = c d I in Z2 Z4 8.49 Now, if Z 1 = Z 3 = Z 5 = Z 2 = Rand Z 4 = 1兾jC, then Z eq becomes Z eq = jCR2 = jLeq Hence, the value of R necessary to yield a 1-H inductance is 1000 . At this point, we must address the question: why go to all this trouble just to make inductance? The answer is size and weight. A 1-H inductor would be very large and heavy. The GIC is easy to construct with integrated circuit components, requires very little space, and weighs only a few grams! 8.10 Design Examples DESIGN EXAMPLE 8.20 In Chapter 4 we found that the op-amp provided us with an easy and effective method of producing controllable voltage gain. From these earlier studies, we have come to expect gain from these “active” devices in a configuration like that shown in Fig. 8.31a. However, irwin08_369-434hr.qxd 28-07-2010 12:03 Page 411 SECTION 8.10 DESIGN EXAMPLES 411 Figure 8.31 vin ± – Active devices used for gain vin R Passive devices used for gain? ± – (a) R Circuit configurations used in Example 8.20. (b) 1/jC + Vin ± – R jL 100 Vo - (c) an experienced engineer has suggested that we could achieve some gain from the proper configuration of “passive” elements as illustrated in Fig. 8.31b, and has proposed the circuit in Fig. 8.31c. Therefore, let us use this suggested configuration in an attempt to design for a gain of 10 at 1 kHz if the load is 100 . The voltage gain of the network in Fig. 8.31c can be expressed as Vo = Vin £ Z Z + 1 § jC where Z = (jL)R jL + R Combining these two equations and rearranging the terms yields the expression Vo = Vin £ jL j c L - 1 L § d + C CR We know that in order to achieve amplification, the denominator must be less than the numerator. In addition, the denominator will be reduced if the reactances of the inductor and capacitor are made equal in magnitude, since they are opposite in sign. Thus, by selecting the parameters such that 2LC = 1, the reactance of the inductor will cancel that of the capacitor. Under this condition, the gain is reduced to Vo = jRC Vin For the given load and frequency values, a capacitor value of 15.9 F will provide the required gain. The inductor value can then be obtained from the constraint 2LC = 1 which yields L = 1.59 mH.It should be noted that if the frequency changes, the impedance of both the inductor and capacitor will also change, thus altering the gain. Finally, we will find in a later chapter that the equation 2LC = 1 is an extremely important expression and one that can have a dramatic effect on circuits. SOLUTION irwin08_369-434hr.qxd 412 28-07-2010 CHAPTER 8 DESIGN EXAMPLE 8.21 SOLUTION 12:03 Page 412 A C S T E A D Y- S TAT E A N A LY S I S A sinusoidal signal, v1(t) = 2.5 cos (t)when added to a dc level of V2 = 2.5 V, provides a 0- to 5-V clock signal used to control a microprocessor. If the oscillation frequency of the signal is to be 1 GHz, let us design the appropriate circuit. As we found in Chapter 4, this application appears to be a natural application for an op-amp summer. However, the frequency of oscillation (1 GHz) is much higher than the maximum frequency most op-amps can handle—typically less than 200 MHz. Since no amplification is required in this case, we should be able to design an op-amp-less summer that, while not precise, should get the job done. Consider the circuit in Fig. 8.32a where inputs v1(t) and V2 are connected to yield the output vo(t). For this application, component A should block any dc component in v1(t) from reaching the output but permit the 1-GHz signal to pass right through. Similarly, component B should pass V2 while blocking any high-frequency signal. Thus, the impedance of component A should be infinite at dc but very low at 1 GHz. And the impedance of component B should be zero at dc but very high at high frequency. Our earlier studies indicate that component A must be a capacitor and component B an inductor. The resulting circuit, called a bias T, is shown in Fig. 8.32b. Figure 8.32 V2 V2 (a) A simple passive summer circuit; (b) a solution—the bias T. L B C A + + + v1(t) vo(t) v1(t) vo(t) - - - - + (a) (b) The values for C and L are dependent on both the signal frequency and the precision required in the summing operation, and can be easily seen by using superposition to investigate the contribution of each input to vo(t). In Fig. 8.33a, the dc voltage V2 has been reduced to zero and an ac circuit has been drawn at a frequency of 1 GHz—that is, the frequency of v1(t). Using voltage division, we can express the output voltage as Vo1 = £ jL jL - j § C V1 = c 2LC 2LC - 1 d V1 8.50 Note that in order to achieve a perfect summer, the voltage division ratio must be unity. However, such a voltage division ratio requires the impractical condition 2LC equal infinity. Instead, we will approach the problem by choosing values for the inductive and capacitive reactances. As stated earlier, the capacitive reactance should be small; we choose 1 . And the inductive reactance should be large—let’s say 10 k. The resulting L and C values are C = 1 1 = = 159 pF XC 2 * 109 and L = XL = 1.59 H Now we consider V2 . In Fig. 8.33b, v1(t) has been reduced to zero and an ac circuit has been drawn at dc—the frequency of V2 . Again, voltage division could be used to express the irwin08_369-434hr.qxd 28-07-2010 12:03 Page 413 SUMMARY Figure 8.33 V2 j(0)L 1/jC + V1 + jL - Vo1 413 + Exploring the bias-T by steady-state ac superposition with (a) V2 = 0 and (b) v1(t) = 0 Vo2 1/j(0)C - (b) (a) output voltage. However, we see that the impedances of the capacitor and inductor are infinity and unity, respectfully. As a result, the output is exactly equal to V2, regardless of the values of C and L! Thus, the output voltage consists of two voltages—one at dc and the other at 1 GHz. The dc component is just V2 = 2.5 V. From Eq. (8.50), the ac component at 1 GHz is Vo1 = c j10,000 d 2.5 / 0° = 2.50025 / 0° V j10,000 - j1 Back in the time domain, the output voltage is, to three significant digits, vo(t) = 2.5 + 2.5 cos C 2A109 BtD V • SUMMARY ■ The sinusoidal function definition The sinusoidal function x(t) = XM sin (t + ) has an amplitude of XM, a radian frequency of , a period of 2兾, and a phase angle of . ■ The phase lead and phase lag definitions If x1(t) = XM1 sin (t + ) and x2(t) = XM2 sin (t + ), x1(t) leads x2(t) by - radians and x2(t) lags x1(t) by - radians. ■ ■ ■ ■ Frequency-domain analysis 1. Represent all voltages, vi(t) , and all currents, ij(t) , as phasors and represent all passive elements by their impedance or admittance. 2. Solve for the unknown phasors in the frequency () domain. 3. Transform the now-known phasors back to the time domain. The phasor definition The sinusoidal voltage v(t) = VM cos (t + ) can be written in exponential form as v(t) = ReCVM ej(t + ) D and in phasor form as V = VM / . The phase relationship in V and i for elements R, L, and C If v and i represent the phase angles of the voltage across and the current through a circuit element, then i = v if the element is a resistor, i lags v by 90° if the element is an inductor, i leads v by 90° if the element is a capacitor. The impedances of R, L, and C Impedance, Z, is defined as the ratio of the phasor voltage, V, to the phasor current, I, where Z=R for a resistor, Z = jL for an inductor, and Z = 1兾jC for a capacitor. The phasor diagrams Phasor diagrams can be used to display the magnitude and phase relationships of various voltages and currents in a network. ■ Solution techniques for ac steady-state problems Ohm’s law KCL and KVL Nodal and loop analysis Superposition and source exchange Thévenin’s theorem Norton’s theorem irwin08_369-434hr.qxd 414 • 28-07-2010 CHAPTER 8 12:03 Page 414 A C S T E A D Y- S TAT E A N A LY S I S PROBLEMS 8.1 Given i(t) 5 cos (400t 120 ) A, determine the period of the current and the frequency in Hertz. 8.7 Find the impedance, Z, shown in Fig. P8.7 at a frequency of 60 Hz. 8.2 Determine the relative phase relationship of the two 2 10 mH waves v1(t) = 10 cos (377t - 30°) V 10 µF 1 Z v2(t) = 10 cos (377t + 90°) V 8.3 Given the following voltage and current: i(t) = 5 sin (377t - 20°) V Figure P8.7 v(t) = 10 cos (377t + 30°) V Determine the phase relationship between i(t) and v(t). 8.4 Determine the phase angles by which v1(t) leads i1(t) and v1(t) leads i2(t), where 8.8 Find the equivalent admittance for the circuit in Fig. P8.8, v1(t) = 4 sin (377t + 25°) V if ◊ = 10 radians/second. i1(t) = 0.05 cos (377t - 20°) A 0.1 H 4 i2(t) = -0.1 sin (377t + 45°) A 8.5 Calculate the current in the capacitor shown in Fig. P8.5 if the voltage input is 0.2 H 0.01 F Yeq (a) v1(t) = 10 cos (377t - 30°) V. (b) v2(t) = 12 sin (377t + 60°) V. 5 8 Give the answers in both the time and frequency domains. Figure P8.8 i(t) + v(t) C=1 F - 8.9 Find the equivalent impedance for the circuit in Fig. P8.9. Figure P8.5 8 –j8 8.6 Find the frequency-domain impedance, Z, as shown in 10 Fig. P8.6. j5 –j2 Zeq Z Figure P8.6 3 j4 j10 4 Figure P8.9 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 415 415 PROBLEMS 8.10 Find the frequency-domain impedance, Z, shown in Fig. P8.10. 8.14 Find the frequency-domain impedance, Z, shown in Fig. P8.14. 2 6 –j1 1 j1 2 1 Z Z –j1 4 j2 j2 j2 j4 –j2 Figure P8.10 Figure P8.14 8.11 Find Z in the network in Fig. P8.11. 2 1 j2 Z 8.15 In the circuit shown in Fig. P8.15, determine the value of the inductance such that the current is in phase with the source voltage. –1j 2 4 j2 2 12 cos (1000t+75°) V Figure P8.11 ± – L 100 F 8.12 Find the impedance, Z, shown in Fig. P8.12 at a Figure P8.15 frequency of 400 Hz. 8.16 Find the value of C in the circuit shown in Fig. P8.16 so that Z is purely resistive at a frequency of 60 Hz. 2 10 mH 1 Z 1 5 mH 10 F C Z Figure P8.12 Figure P8.16 8.13 Find the impedance, Z, shown in Fig. P8.13 at a frequency of 60 Hz. 2 8.17 Find the value of the capacitance, C, shown in the circuit in Fig. P8.17 so that i(t) will be in phase with the source voltage. 4 i(t) 8 Z Figure P8.13 10 mH 500 F v(t) = 60 cos (250t+30°) V Figure P8.17 ± – 15 C 40 mH irwin08_369-434hr.qxd 416 28-07-2010 CHAPTER 12:03 8 Page 416 A C S T E A D Y- S TAT E A N A LY S I S 8.18 The impedance of the network in Fig. P8.18 is found to be purely real at f = 400 Hz. What is the value of C? 8.23 In the diagram in Fig. P8.23, v(t)=50 cos(10t+10°) V and i(t)=25 cos(10t+41°) A. Is the impedance of the BOX inductive or capacitive? Explain your answer. i(t) 6 Z C v(t) + – 10 mH Box Figure P8.23 Figure P8.18 8.19 Find the frequency at which the circuit shown in Fig. P8.19 is purely resistive. 8.24 Determine the value of in Fig. 8.24 such that the peak value of i(t) is 2 A. i(t) 1 Z 5 mH 10 1 mF v(t) = 25 cos t V + – 0.002 F Figure P8.19 8.20 At = 100 rad/s, the equivalent admittance for the Figure P8.24 circuit in Fig. P8.20 is 0.1 + j0.2 S. Calculate values for R and C R 8.25 In the diagram in Fig. P8.25 when v(t)=5 cos 500t V, i(t)=0.4 cos(500t-30°) A. Calculate i(t) if v(t)= 5 cos 1000t V. Y C i(t) v(t) + – Figure P8.20 Linear circuit 8.21 The admittance of the box in Fig. P8.21 is 0.1 + j0.2 S at 500 rad/s. What is the impedance at 300 rad/s? Figure P8.25 Y 8.26 Draw the frequency-domain network and calculate vo(t) in the circuit shown in Fig. P8.26 if i1(t) is 200 cos (105t + 60°) mA, i2(t) is 100 sin(105t + 90°) mA, and vS(t) = 10 sin(105t) V. Also, use a phasor diagram to determine vC(t). Figure P8.21 8.22 The impedance of the box in Fig. P8.22 is 5 + j4 at 1000 rad/s. What is the impedance at 1300 rad/s? - vC (t) + Z i1(t) i2(t) 30 vo(t) - Figure P8.22 Figure P8.26 + 250 nF ± vS(t) – irwin08_369-434hr.qxd 28-07-2010 12:03 Page 417 417 PROBLEMS 8.27 Find vx(t) and vR(t) in the circuit in Fig. P8.27 if v(t)=50 cos10t V. 3 8.29 Find io(t) in the circuit in Fig. P8.29 if v(t)=50 cos 100t V. 4 0.2 H + vR(t) + – vx(t) 20 5 2000 µF 5 2H 6 20 mH 0.01 F v(t) + – 1000 µF io(t) 3 + v(t) – 50 mH – Figure P8.29 Figure P8.27 8.30 Find v0(t) and io(t) in Fig. P8.30. 8.28 Calculate vC(t) in Fig. P8.28. 3H 4 0.05 H 5 8 2H 7 + – 100 cos 100t V 1000 µF + 10 0.04 H 0.05 F vC (t) + – + vo(t) 25 cos 2t V io(t) 5 – – Figure P8.30 Figure P8.28 8.31 Calculate v1(t) if i1(t)=7 cos 100t A and i2(t)=3cos(100t+45º) in Fig. P8.31. + i1(t) 10 v2(t) i2(t) 50 mH 1000 F – Figure P8.31 8.32 Calculate vo(t) in Fig. P8.32. 2 2H 6 50 cos 5t V 0.05 F + vo(t) – Figure P8.32 + – 4 1H 0.2 F 5 3 0.1 F 5 6 irwin08_369-434hr.qxd 418 28-07-2010 CHAPTER 12:03 8 Page 418 A C S T E A D Y- S TAT E A N A LY S I S 8.33 Calculate vx(t) in the circuit in Fig. 8.33 if v(t)=50 cos(10t-30°) V. 2 2 0.025 F + v(t) + – vx(t) 0.025 F 0.2 H 5 0.8 H – Figure P8.33 8.34 Calculate i1(t), i2(t), and vx(t) in Fig. P8.34. 4 + i1(t) 5 50 cos 10t V + – 10 0.05 F vx(t) 0.4 H 0.025 F 0.5 H – i2(t) Figure P8.34 8.35 Calculate i1(t) and i2(t) in Fig. P8.35. 4 0.025 F i2(t) + vx(t) 50 cos 10t V + – i1(t) 5 – 10 0.05 F 0.4 H 0.5 H Figure P8.35 8.36 Find vx(t) in the circuit in Fig. P8.36. 10 0.005 F 0.1 H + 100 cos 40t V ± – ± – vx(t) 5 Figure P8.36 0.2 H 40 cos (40t-30°) V 0.75 vx(t) ± – irwin08_369-434hr.qxd 28-07-2010 12:03 Page 419 419 PROBLEMS 8.37 Find vo(t) in the network in Fig. P8.37. 10 50 cos 25t V 0.008 F 0.1 H vo(t) ± – ± – 2vx(t) + vx(t) 0.2 H 5 Figure P8.37 8.38 Find v1(t) and v2(t) in the circuit in Fig. P8.38. 12 cos(10t-25°) V 2 + - + v2(t) 4 cos10t A 0.2 H v1(t) 0.02 F 2 cos(10t+15°) A - Figure P8.38 8.42 Find Vo in the network in Fig. P8.42. 8.39 Find the voltage V shown in Fig. P8.39. 1 –j1 2 + 100 0° V ± – –j1 + V j1 12 0° V + 4 2 Vo - - Figure P8.39 Figure P8.42 8.40 Find the voltage Vo shown in Fig. P8.40. 2 8.43 Find Vo in the network in Fig. P8.43. j10 + + 10 30° V ± – –j12 Vo Vo 2 60° A 8.41 Find the frequency-domain voltage Vo shown in Fig. P8.41. 8.44 Given the network in Fig. P8.44, determine the value of Vo if VS = 24 / 0°V. 1 –j12 j2 2 + Vo + + VS ± – – j1 - Figure P8.41 100 30° V Figure P8.43 Figure P8.40 5 30° A ± – j40 - - 15 30 V1 - Figure P8.44 2 Vo - irwin08_369-434hr2.qxd 420 2-08-2010 CHAPTER 16:44 8 Page 420 A C S T E A D Y- S TAT E A N A LY S I S 8.45 If V1 = 4 / 0° V , find I o in Fig. P8.45. 8.49 If I o = 4 / 0° A in the circuit in Fig. P8.49, find I x. –j2 ⍀ - ± – 1⍀ + V1 1⍀ 2⍀ j1 ⍀ IS 1⍀ Ix Io –j1 ⍀ j1 ⍀ 1⍀ 1⍀ 2 0° A 12 0° V ± – 1⍀ 4 0° V Io Figure P8.49 Figure P8.45 8.50 If I o = 4 / 0° A in the network in Fig. P8.50, find I x. 8.46 Find VS in the network in Fig. P8.46, if V1 = 4 / 0° V . –j1 ⍀ + + 1⍀ V1 Ix j1 ⍀ VS j1 ⍀ – j1 ⍀ 2⍀ 1⍀ 1⍀ 2⍀ - R2 Figure P8.46 2 0° A 1⍀ –j1 ⍀ 12 0° V 1⍀ Io 8.47 In the network in Fig. P8.47, Vo is known to be 4 / 45° V. Find Z. 2⍀ 12 0° V 1⍀ ± – 1⍀ 1⍀ V1 V2 1⍀ 1⍀ 1⍀ Z 8.51 Using nodal analysis, find I o in the circuit in Fig. P8.51. + – j1 ⍀ ± – Figure P8.50 Vo - ± – –j1 ⍀ 1⍀ 12 0° V 2 0° A 4 0° A Io Figure P8.47 Figure P8.51 8.48 In the network in Fig. P8.48, I o = 4 / 0° A . Find I x. 8.52 Use nodal analysis to find I o in the circuit in Fig. P8.52. V j1 ⍀ 1⍀ –j1 ⍀ –j1 ⍀ 2⍀ 1⍀ j2 ⍀ 4 0° A 2 0° A 1⍀ Ix Figure P8.48 ± – 12 0° V 1⍀ 12 0° V ± – Io Figure P8.52 ± – 2⍀ 6 0° V Io irwin08_369-434hr.qxd 28-07-2010 12:03 Page 421 PROBLEMS 421 8.53 Find Vo in the network in Fig. P8.53 using nodal analysis. V 2 1 –j1 j2 ± – 12 0° V 2 0° A + 2 ± – 4 0° V Vo - Figure P8.53 8.54 Use the supernode technique to find I o in the circuit in Fig. P8.54. 12 0° V –j1 2 + 1 – j2 j2 2 Io Figure P8.54 8.55 Use nodal analysis to find Vo in the circuit in Fig. P8.55. 6 0° V 12 0° V ±– –± –j1 1 1 + j2 1 Vo 2 0° A - Figure P8.55 8.56 Find I o in the circuit in Fig. P8.56 using nodal analysis. 8.57 Use nodal analysis to find I o in the circuit in Fig. P8.57. Io 2 1 2 –j2 2 0° A ± – 1 6 0° V j1 4 0° A 12 0° V ±– 12 0° V ± – j1 ± – 6 0° V 2 0° A –j1 1 1 Io Figure P8.56 Figure P8.57 1 irwin08_369-434hr.qxd 422 28-07-2010 CHAPTER 12:03 8 Page 422 A C S T E A D Y- S TAT E A N A LY S I S 8.58 Use nodal analysis to find Vo in the circuit in Fig. P8.58. 8.62 Use nodal analysis to find Vo in the network in Fig. 8.62. + 4 Vo + – 1 + – + –j –j1 1 Vx 1 1 + 1 – 4V 2 0° A Vo 6 0° V Vo 1 – 1 + – + – 4 Vx – Figure P8.62 Figure P8.58 8.59 Use nodal analysis to find Vo in the network in Fig. P8.59. 8.63 Use nodal analysis to find Vo in the circuit in Fig. P8.63. + 4V + – Ix j –j1 4 Ix 1 1 + Vo + + – 1 4 Vo Vx – Vo – 1 1 1 + – 1 6 0° V + – 2 Vx – Figure P8.59 Figure P8.63 8.60 Use nodal analysis to find Vo in the circuit in Fig. P8.60. 8.64 Use nodal analysis to find Vo in the circuit in Fig. P8.64. –j1 + 1 4V + – + Vx 1 + – + 1 Vx 4 Vx 6 0° V 2 0° A – + – 1 4 0° A – + 1 –j1 Vo – + – 1 4 Vx – Figure P8.60 8.61 Use nodal analysis to find Vo in the circuit in Fig. P8.61. 4 Ix Figure P8.64 8.65 Use nodal analysis to find I o in the circuit in Fig. P8.65. + Ix –j1 + 6 0° V 1 + –j1 Vx 1 1 Ix Figure P8.61 Vo 1 Io – 1 1 2 0° A Vo 4 Ix – Figure P8.65 + – 4 0° V + – 2 Vx irwin08_369-434hr.qxd 28-07-2010 12:03 Page 423 423 PROBLEMS 8.66 Use nodal analysis to find Vx in the circuit in Fig. P8.66. 8.70 Use loop analysis to find Vo in the circuit in Fig. P8.70. 12 0° V + ±– j1 + 1 2 Vx –j1 Vx 1 + 1 Vo - 1 1 Vo 4 0° A 6 0° V Figure P8.66 + – –j1 – Figure P8.70 8.67 Find the voltage across the inductor in the circuit shown in Fig. P8.67 using nodal analysis. 8.71 Use loop analysis, to find Vo in the circuit in Fig. P8.71. V1 10 30° V ± – 4 –j2 V2 + 1 2Ix j1 Ix 1 4 0° A 1 1 Figure P8.67 6 0° V + – Vo –j1 – 8.68 Use mesh analysis to find Vo in the circuit shown in Fig. P8.68. 8.72 Using loop analysis, find I o in the network in Fig. P8.72. 12 45° V –j1 12 0°V –± 6 0° V Figure P8.71 + j2 – ± I1 I2 2 2 + Vo 2 0° A - j1 Figure P8.68 –j2 2 Io 8.69 Use mesh analysis to find Vo in the circuit shown in Fig. P8.69. Figure P8.72 8.73 Find Vo in the circuit in Fig. P8.73 using mesh analysis. j2 4 j2 2 –j4 12 0° V 4 0° A ± – + I2 2 0° A I1 + 2 ±– 1 Vo 6 0° V 4 0° A Figure P8.69 2 – j1 4 90° A Figure P8.73 Vo - irwin08_369-434hr.qxd 424 28-07-2010 CHAPTER 12:03 8 Page 424 A C S T E A D Y- S TAT E A N A LY S I S 8.74 Use mesh analysis to find Vo in the circuit in Fig. P8.74. 8.78 Find Vo in the network in Fig. P8.78. 4 0° A j1 6 0° A –j1 4 0°A –j2 + j1 2Ix 2 Vo 1 Vo 2 0° A - - Figure P8.78 Figure P8.74 8.79 Use loop analysis to find Vo in the network in Fig. P8.79. 8.75 Use loop analysis to find I o in the network in Fig. P8.75. + Vx Io + – 1 1 ± – + 1 –j1 + Vx – 1 4 0° A 1 2 0° A 6 0° V – 4 30° A – j1 1 2Vx Ix 1 + 2 1 + 1 2 0° A Vo – Vx Figure P8.79 - 8.80 Use loop analysis to find Vo in the circuit in Fig. P8.80. Figure P8.75 + 8.76 Find Vo in the network in Fig. P8.76. –j1 4 Ix j1 ± – 12 0° V 2 + 4V – Figure P8.80 2 0° A 8.81 Use loop analysis to find Vo in the network in Fig. P8.81. Ix Figure P8.76 8.77 Determine Vo in the circuit in Fig. P8.77. 2 –j1 ± – j2 12 0° V –j1 1 2 + Figure P8.77 Vo Ix Vo –j1 - 6 0° A + – 1 2 16 0° V 1 1 + 1 Vo 1 6V + – - + 1 Vo – Figure P8.81 4 Ix irwin08_369-434hr.qxd 28-07-2010 12:03 Page 425 425 PROBLEMS 8.82 Use loop analysis to find Vo in the circuit in Fig. P8.82. 8.88 Find Vo in the network in Fig. P8.88 using superposition. 2 + 1 4 Ix 2 0° A 1 1 + – –j1 12 0° V –± Vo –j2 + j4 2 Vo 6V - – Ix Figure P8.88 Figure P8.82 8.89 Find Vo in the network in Fig. P8.89 using superposition. 8.83 Use loop analysis to find Vo in the circuit in Fig. P8.59. 8.85 Use superposition to find Vo in the network in Fig. P8.85. 16 0° V + - 2 – j2 12 0° V –± 1 2 Vo + 1 – j1 4 0° A + 1 8.84 Use loop analysis to find Vo in the network in Fig. P8.63. j3 4 0° A Vo - - Figure P8.85 Figure P8.89 8.90 Use superposition to find Vo in the circuit in Fig. P8.90. 8.86 Use superposition to determine Vo in the circuit in Fig. 8.86. 1 j1 2 0° A j1 –j1 1 + 6 0° V + 4 0° A 1 6 0° A + – –j1 Vo + 1 1 6 0° V Vo – - Figure P8.90 Figure P8.86 8.87 Using superposition, find Vo in the circuit in Fig. P8.87. 1 + –j1 ± – 2 0° A + Vo 2 + - 1 j2 6 0° V 8.91 Use superposition to find Vo in the network in Fig. 8.91. 1 1 –j1 Vo 4 0° A – - Figure P8.87 6 0° V Figure P8.91 irwin08_369-434hr2.qxd 426 2-08-2010 CHAPTER 16:44 8 Page 426 A C S T E A D Y- S TAT E A N A LY S I S 8.92 Use superposition to find Vo in the circuit in Fig. P8.92. 8.94 Use source transformation to determine I o in the network in Fig. P8.94. + 1⍀ 4 0° A Vo 1⍀ V2 1⍀ – ± 1⍀ –j1 ⍀ 1⍀ 12 0° V + – 6 0° V V1 1⍀ 2 0° A 4 0° A Io –j1 ⍀ – Figure P8.94 8.95 Use source exchange to determine Vo in the network in Figure P8.92 Fig. P8.95. 8.93 Use source exchange to find the current I o in the network in Fig. P8.93. –± 2⍀ –j1 ⍀ 1⍀ 12 0° V 12 0° V + - + - + 6 0° V 1⍀ 2⍀ 2 0° A - –j1 ⍀ 2⍀ 2 0° A Io 4 0° A Figure P8.95 Figure P8.93 8.96 Use source transformation to determine I o in the network in Fig. P8.96. V 2⍀ –j1 ⍀ ± – 2⍀ j2 ⍀ 4 0° A 12 0° V ± – 6 0° V Io Figure P8.96 8.97 Use source transformation to find Vo in the circuit in Fig. P8.97. 8.98 Use source transformation to find Vo in the circuit in Fig. P8.98. – j2 ⍀ + 12 0° V + - –j2 ⍀ Vo + 5⍀ 2 0° A 1⍀ 1⍀ 12 ⍀ Vo Vo - 1⍀ 1⍀ 2 0° R 4 0° V + - Figure P8.97 2⍀ 2⍀ Figure P8.98 + - 6 0° V irwin08_369-434hr.qxd 28-07-2010 12:03 Page 427 427 PROBLEMS 8.99 Use source transformation to find I o in the circuit in Fig. P8.99. 3 6 0° V Io 1 + - 12 4 0° A 2 0° A + - 6 8 0° V 4 –j3 Figure P8.99 8.100 Use source transformation to find I o in the network in 8.102 Apply Thévenin’s theorem twice to find Vo in the cir- Fig. P8.100. cuit in Fig. P8.102. 1 2 2 1 + 6 0° V + - –j1 j1 1 Vo - 4 0° A + - 2 2 10 0° V Figure P8.102 8.103 Use Thévenin’s theorem to find I o in the network in –j1 Fig. P8.103. 2 0° A 3 Io Io 2 4 Figure P8.100 12 0° V 8.101 Use Thevenin’s theorem to find Vo in the circuit in + - – j2 2 4 0° A Figure P8.103 Fig. P8.101. 8.104 Use Thévenin’s theorem to determine I o in the circuit j2 4 in Fig. P8.104. 2 0° A –j4 12 0° V ± – I1 I2 + 2 4 90° A + Vo 1 - Figure P8.101 –j2 6 0° V 1 Io Figure P8.104 1 4 0° A irwin08_369-434hr.qxd 428 28-07-2010 CHAPTER 8 12:03 Page 428 A C S T E A D Y- S TAT E A N A LY S I S 8.105 Use Thévenin’s theorem to find Vo in the circuit 8.106 Use Thévenin’s theorem to find Vo in the network in in Fig. P8.105. Fig. P8.106. + + - 1 1 1 6 0° V Vo 1 1 –j1 + 1 4 0° A 2 6 0° V 4 0° A + - Vo – j1 - - Figure P8.105 Figure P8.106 8.107 Use Thévenin’s theorem to find Vo in the network in Fig. P8.107. 1 + 1 4 0° V + - + 6 0° V –j1 2 0° A 1 Vo - Figure P8.107 8.108 Use Thévenin’s theorem to find I o in the network in Fig. P8.108. 6 0° V + 1 4 0° V + - –j1 1 1 2 0° A 1 2 0° A Io Figure P8.108 8.109 Use Thévenin’s theorem to find Vo in the network in Fig. P8.109. 8.110 Given the network in Fig. P8.110, find the Thévenin’s equivalent of the network at terminals A–B. 1 V2 –j1 12 0° V –j1 6 0° A 1 + 4 0° A ±– 1 + 4 0° A j2 2 Vo V1 V3 j1 - ± – 1 6 0° V A 4 0° V B Figure P8.109 Figure P8.110 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 429 429 PROBLEMS 8.111 Find Vo in the network in Fig. P8.111 using Thévenin’s theorem. Fig. P8.105. 8.118 Use Norton’s theorem to find Vo in the circuit in 12 0° V Fig. P8.107. ±– + –j1 1 2 Vx 8.117 Use Norton’s theorem to find Vo in the network in j1 + 1 Vx - Vo - 8.119 Use Norton’s theorem to find I o in the circuit in Fig. P8.108. 8.120 Use Norton’s theorem to find Vo in the network in Fig. P8.120. Figure P8.111 8.112 Find the Thévenin’s equivalent for the network in + Fig. P8.112 at terminals A–B. – j1 4 0° A A + –j1 4 0° A 1 + Vx - 1 1 Vx - 2Vx j1 1 Vo - j1 2 Vx Figure P8.120 B 8.121 Find Vo using Norton’s theorem for the circuit in Fig. P8.121. Figure P8.112 8.113 Find Vx in the circuit in Fig. P8.113 using Norton’s theorem. 4 0° V ± – 11.3 45° V 1 –± j4 2 0° A + –j3 Vx Vx 8 0° V + 1 + 10 ± – –j1 ± – j1 2Vx 1 Vo - - Figure P8.121 Figure P8.113 8.114 Find I o in the network in Fig. P8.114 using Norton’s theorem. 8.122 Use Norton’s theorem to find Vo in the circuit in Fig. P8.122. –j2 –j1 4 Ix j1 6 45° A 2 2 0° A 1 Io Figure P8.114 8.116 Use Norton’s theorem to find I o in the circuit in Fig. P8.104. + 1 4 0° A 1 2 Ix 8.115 Use Norton’s theorem to find I o in the circuit in Fig. P8.103. 2 0° A Figure P8.122 8.123 Use Norton’s theorem to find Vo in the circuit in Fig. P8.61. Vo – irwin08_369-434hr.qxd 430 28-07-2010 CHAPTER 12:03 8 Page 430 A C S T E A D Y- S TAT E A N A LY S I S 8.124 Use source transformation to find Vo in the circuit in Fig. P8.124. Solve the remaining problems using computational methods. 8.128 Apply Norton’s theorem to find Vo in the network in Fig. P.8.128. –j1 + 12 0° V + – –j2 6 0° V 6 ± – 1 6 Vo 1 j1 12 + 1 Vo 2 0° A 1 4 0° A - Figure P8.124 Figure P8.128 8.125 Calculate the Thévenin equivalent impedance Z Th in the 8.129 Find Vo in the circuit in Fig. P8.129. circuit shown in Fig. P8.125. 6 0° V + 4 Ix ZTh Ix ± – 2 0° A 12 0° V –j1 –j3 3 1 1 ±– + 1 1 Vo Figure P8.125 j1 1 - 8.126 Find the Thévenin equivalent for the network in Figure P8.129 8.130 Find the node voltages in the network in Fig. P8.130. Fig. P8.126 at terminals A–B. 1 Ix 1 A – j1 2 –j1 2 0° A ± – 2Ix j1 B 12 0° V 1 j1 1 ± – –j2 2 Figure P8.126 8.127 Find the Thévenin equivalent of the network in Fig. P8.127 at terminals A–B. + Vx Figure P8.130 8.131 Determine Vo in the network in Fig. P8.131. - 1 1 A 1 + j1 – ± –j1 2Vx + 1 1 j1 12 30° V 1 Vo - B Figure P8.127 ± – 1 6 0° V –j1 3 2 0° A Figure P8.131 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 431 431 PROBLEMS 8.132 Use Thévenin’s theorem, to determine I o in the network 8.136 If the network in Fig. P8.136 operates at f = 400 Hz , in Fig. P8.132. find the current I o . Io 1 1 4 0° A ± – 12 0° V 1 ± – 12 0° V 2 Io –j1 1 1 ± – j1 1 2 –j2 1 6 0° V Figure P8.132 ± – 2 0° A j1 6 0° V Figure P8.136 8.133 Find I o in the network in Fig. P8.133. 8.137 The network in Fig. P8.137 operates at f = 60 Hz . Find the voltage Vo . 1 2 0° A 1 1 –j1 4 0° A 1 1 2 0° A 2 Io ± – 12 0° V ± – j1 1 –j1 2 + 6 0° V ± – j2 1 12 0° V Vo - Figure P8.133 Figure P8.137 8.134 Use both nodal analysis and loop analysis to find I o in the network in Fig. P8.134. 1 – j1 2 0° A 1 Io + 2Vx 8.138 Find I o in the network in Fig. P8.138. ± – 1 2 Ix 1 –± – j1 1 6 0° V 12 0° V ±– –± 1 Io 1 12 0° V j1 1 Vx - 1 1 2 0° A j1 4 0° A Ix Figure P8.134 Figure P8.138 8.135 Given the circuit in Fig. P8.135, at what frequency are the magnitudes of iC(t) and iL(t) equal? 8.139 Find I o in the network in Fig. P8.139. R1 1 100 R2 vin(t) ± 5 cos(t) V – 2Ix 1 100 mH C Io 1 L iC(t) 100 F Figure P8.135 Ix j1 150 1 1 1 iL(t) Figure P8.139 ± – 6 0° V –j1 + 4 0° V irwin08_369-434hr.qxd 432 28-07-2010 CHAPTER 12:03 8 Page 432 A C S T E A D Y- S TAT E A N A LY S I S 8.141 Find I o in the circuit in Fig. P8.141. 8.140 Determine I o in the network in Fig. P8.140. 1 1 j1 –j1 1 1 12 0° V – j1 ± – 12 0° V + - Io 2Vx 1 Io 1 6 0° V 1 2Ix Figure P8.141 Figure P8.140 8.142 Find vo(t) in the network in Fig. P8.142. 50F ± – 2ix 2 mH 4 k ix + vo(t) - Figure P8.142 8.143 The network in Fig. P8.143 operates at 60 Hz. Find the currents I o and I x. Ix 10 Io j18.85 2 Ix –j5.305 ± – 12 0° V Figure P8.143 8.144 Find vo(t) in the network in Fig. P8.144. 50F 16 cos (377t+45°) V Ix 1 - 16 cos (377t+45°) V 2 0° A + Vx 1 2 0° A 1 + 1 ± – 2 ix 2 mH 4 k ix 8.145 The network in Fig. P8.145 operates at 60 Hz. Find the currents I o and I x. Ix Io Figure P8.145 j18.85 2 Ix vo(t) - Figure P8.144 10 + –j5.305 ± – 12 0° V ± – j1 irwin08_369-434hr.qxd 28-07-2010 12:03 Page 433 TYPICAL PROBLEMS FOUND ON THE FE EXAM • TYPICAL PROBLEMS FOUND ON THE FE EXAM 8PFE-1 Find Vo in the network in Fig. 8PFE-1. a. b. c. d. 4.62⬔30.4° V 7.16⬔-26.6° V 3.02⬔24.3° V 5.06⬔-71.6° V –± j1 + 12 0° V – j1 2 1 Vo 2 0° A - Figure 8PFE-1 8PFE-2 Find Vo in the circuit in Fig. 8PFE-2. a. b. c. d. 25.4⬔10.25° V 20.1⬔4.63° V 30.8⬔8.97° V 18.3⬔12.32° V 2 Ix 1 2 + –j1 ± – 12 0° V Vo 2 0° A - Ix Figure 8PFE-2 8PFE-3 Find Vo in the network in Fig. 8PFE-3. a. b. c. d. 8.24⬔-30.96° V 2.06⬔20.84° V 16.96⬔45° V 10.42⬔30° V 2 –j1 + j2 12 0° V ± – 4 ± – 2 Vo Figure 8PFE-3 Vo 433 irwin08_369-434hr.qxd 434 28-07-2010 CHAPTER 8 12:03 Page 434 A C S T E A D Y- S TAT E A N A LY S I S 8PFE-4 Determine the midband (where the coupling capacitors can be ignored) gain of the single-stage transistor amplifier shown in Fig. 8PFE-4. a. 110.25 b. 133.33 c. 26.67 d. 95.75 + 1 k ± V S – 5 k 40*10 Vx –3 + Vx 12 k 6 k - - Figure 8PFE-4 8PFE-5 What is the current Io in the circuit in Fig. 8PFE-5? a. b. c. d. 6.32⬔30.31° A 2.75⬔21.43° A 1.48⬔32.92° A 5.23⬔40.15° A 3 –j1 j3 6 0° V Vo ± – 1 1 Figure 8PFE-5 Io irwin09_435-490hr.qxd 28-07-2010 12:00 Page 435 C H A PT E R STEADY-STATE POWER ANALYSIS 9 THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Know how to calculate instantaneous and average power in ac circuits ■ Be able to calculate the maximum average power transfer for a load in an ac circuit ■ Know how to calculate the effective or rms value for a periodic waveform ■ Know how to calculate real power, reactive power, complex power, and power factor in ac circuits ■ Understand how to correct the power factor in ac circuits Mlenny Photography/Alexander Hafemann/ iStockphoto S Solar mirror arrays Enough solar energy strikes Earth every ■ Understand the importance of safety and the consequences of ignoring it when working with power than other solar technology. Aligning this mirror’s direction day to power our homes and businesses for almost 30 years. using feedback control can maximize the effective sun’s rays However, today’s solar power systems are able to capture only seen by the mirrors, even on cloudy days. a miniscule fraction of that energy. The ongoing challenge in This chapter is aimed at filling a gap in your study of the solar power industry is to develop collection technologies circuit analysis: finding the average power supplied or that convert more of the sun’s energy into usable power. absorbed by each element in ac circuits. This average real Three technologies that strive to capture solar power power is the product of effective values of sinusoidal voltages more efficiently are parabolic troughs, tower systems, and and currents—as expected—but also a power factor that dish/engine systems. Trough systems use mirrors to multiply depends on their phase differences. Leading and lagging the sun’s energy from 30 to 60 times its normal intensity power factors are discussed, and complex power is intro- to a receiver pipe containing synthetic oil at the focal line. duced as a common term in power distribution. Whatever the Power towers focus thousands of mirrors onto a receiver at generation source, ac power can be delivered more efficiently the top of the tower. Solar Two near Barstow, CA is the by improving the power factor. Getting a handle on power world’s largest solar power tower. Dish/engine systems use a factors is far from an academic exercise; it is central to under- mirrored array to give 30% conversion efficiencies—higher standing how to wring more power out of the sun. 435 irwin09_435-490hr.qxd 436 28-07-2010 CHAPTER 9 12:00 Page 436 S T E A D Y- S TAT E P O W E R A N A LY S I S 9.1 By employing the sign convention adopted in the earlier chapters, we can compute the instantaneous power supplied or absorbed by any device as the product of the instantaneous voltage across the device and the instantaneous current through it. Consider the circuit shown in Fig. 9.1. In general, the steady-state voltage and current for the network can be written as Instantaneous Power v(t) = VM cos At + v B 9.1 i(t) = IM cos At + i B i(t) 9.2 The instantaneous power is then v(t) ± – Z p(t) = v(t)i(t) = VM IM cos At + v B cos At + i B 9.3 Employing the following trigonometric identity, Figure 9.1 Simple ac network. cos 1 cos 2 = 1 C cos A1 - 2 B + cos A1 + 2 B D 2 9.4 we find that the instantaneous power can be written as p(t) = VM IM C cos Av - i B + cos A2t + v + i B D 2 9.5 Note that the instantaneous power consists of two terms. The first term is a constant (i.e., it is time independent), and the second term is a cosine wave of twice the excitation frequency. We will examine this equation in more detail in Section 9.2. EXAMPLE 9.1 SOLUTION The circuit in Fig. 9.1 has the following parameters: v(t) = 4 cos (t + 60°) V and Z = 2 / 30° ⍀. We wish to determine equations for the current and the instantaneous power as a function of time and plot these functions with the voltage on a single graph for comparison. Since I = 4 / 60° 2 / 30° = 2 / 30° A [hint] Note that p(t) contains a dc term and a cosine wave with twice the frequency of v(t) and i(t). then i(t) = 2 cos (t + 30°) A From Eq. (9.5), p(t) = 4[cos(30°) + cos(2t + 90°)] = 3.46 + 4 cos(2t + 90°) W A plot of this function, together with plots of the voltage and current, is shown in Fig. 9.2. As can be seen in this figure, the instantaneous power has a dc or constant term and a second term whose frequency is twice that of the voltage or current. irwin09_435-490hr.qxd 28-07-2010 12:00 Page 437 SECTION 9.2 8.0 AV E R AG E P OW E R Figure 9.2 p(t) 6.0 437 Plots of v(t), i(t), and p(t) for the circuit in Example 9.1 using f ⫽ 60 Hz. 4.0 2.0 0.0 –2.0 –4.0 –6.0 i(t) v(t) t(s) –8.0 0.000 0.003 0.006 0.009 0.012 0.015 0.018 0.021 0.024 0.027 0.030 9.2 The average value of any periodic waveform (e.g., a sinusoidal function) can be computed by integrating the function over a complete period and dividing this result by the period. Therefore, if the voltage and current are given by Eqs. (9.1) and (9.2), respectively, the average power is Average Power t0 + T 1 P = p(t) dt T 3t0 = t0 + T 1 VM IM cos At + v B cos At + i B dt T 3t0 9.6 where t0 is arbitrary, T = 2兾 is the period of the voltage or current, and P is measured in watts. Actually, we may average the waveform over any integral number of periods so that Eq. (9.6) can also be written as P = t0 + nT 1 VM IM cos (t + v) cos (t + i) dt nT 3t0 9.7 where n is a positive integer. Employing Eq. (9.5) for the expression in (9.6), we obtain P = t0 + T VM IM 1 C cos (v - i) + cos (2t + v + i)D dt T 3t0 2 9.8 We could, of course, plod through the indicated integration; however, with a little forethought we can determine the result by inspection. The first term is independent of t, and therefore a constant in the integration. Integrating the constant over the period and dividing by the period simply results in the original constant. The second term is a cosine wave. It is well known that the average value of a cosine wave over one complete period or an integral number of periods is zero, and therefore the second term in Eq. (9.8) vanishes. In view of this discussion, Eq. (9.8) reduces to P = 1 V I cos Av - i B 2 M M 9.9 [hint] Note that since cos (-) = cos (), the argument for the cosine function can be either v - i or i - v . In addition, note that v - i is the angle of the circuit impedance as shown in Fig. 9.1. Therefore, for a purely resistive circuit, P = 1 V I 2 M M 9.10 A frequently used equation for calculating the average power. irwin09_435-490hr.qxd 438 28-07-2010 CHAPTER 9 12:00 Page 438 S T E A D Y- S TAT E P O W E R A N A LY S I S and for a purely reactive circuit, 1 VM IM cos(90°) 2 = 0 P = Because purely reactive impedances absorb no average power, they are often called lossless elements. The purely reactive network operates in a mode in which it stores energy over one part of the period and releases it over another. EXAMPLE 9.2 We wish to determine the average power absorbed by the impedance shown in Fig. 9.3. I Figure 9.3 Example RL circuit. 2⍀ ± – 10 60° V j2 ⍀ SOLUTION From the figure we note that I = VM / v 10 / 60° V = = = 3.53 / 15° A Z 2 + j2 2.83 / 45° Therefore, IM = 3.53 A and i = 15° Hence, 1 V I cos (v - i) 2 M M 1 = (10)(3.53) cos (60° - 15°) 2 P = = 12.5 W Since the inductor absorbs no power, we can employ Eq. (9.10), provided that VM in that equation is the voltage across the resistor. Using voltage division, we obtain VR = A10 / 60°B(2) 2 + j2 = 7.07 / 15° V and therefore, 1 (7.07)(3.53) 2 = 12.5 W P = In addition, using Ohm’s law, we could also employ the expressions P = 1 V2M 2 R P = 1 2 IM R 2 or where once again we must be careful that the VM and IM in these equations refer to the voltage across the resistor and the current through it, respectively. irwin09_435-490hr.qxd 28-07-2010 12:00 Page 439 SECTION 9.2 For the circuit shown in Fig. 9.4, we wish to determine both the total average power absorbed and the total average power supplied. I V I2 I1 ± – 12 45°V AV E R AG E P OW E R EXAMPLE 9.3 Figure 9.4 2⍀ Example circuit for illustrating a power balance. 4⍀ –j1 ⍀ SOLUTION From the figure we note that I1 = I2 = 12 / 45° 4 = 3 / 45° A 12 / 45° 2 - j1 12 / 45° = 2.24 / -26.57° = 5.36 / 71.57° A and therefore, I = I1 + I2 = 3 / 45° + 5.36 / 71.57° = 8.15 / 62.10° A The average power absorbed in the 4-⍀ resistor is P4 = 1 1 V I = (12)(3) = 18 W 2 M M 2 The average power absorbed in the 2-⍀ resistor is P2 = 1 2 1 I M R = (5.34)2(2) = 28.7 W 2 2 Therefore, the total average power absorbed is PA = 18 + 28.7 = 46.7 W Note that we could have calculated the power absorbed in the 2-⍀ resistor using 1兾2 V2M兾R if we had first calculated the voltage across the 2-⍀ resistor. The total average power supplied by the source is 1 V I cos (v - i) 2 M M 1 = (12)(8.15) cos (45° - 62.10°) 2 = 46.7 W PS = Thus, the total average power supplied is, of course, equal to the total average power absorbed. 439 irwin09_435-490hr.qxd 440 28-07-2010 CHAPTER 9 12:00 Page 440 S T E A D Y- S TAT E P O W E R A N A LY S I S Learning Assessments E9.1 Find the average power absorbed by each resistor in the network in Fig. E9.1. ANSWER: P2⍀ = 7.20 W; P4⍀ = 7.20 W. 2⍀ 12 60° V ± – 4⍀ –j4 ⍀ Figure E9.1 E9.2 Given the network in Fig. E9.2, find the average power absorbed by each passive circuit element and the total average power supplied by the current source. ANSWER: P3⍀ = 56.60 W; P4⍀ = 33.96 W; PL = 0; PCS = 90.50 W. 4⍀ 3⍀ j2 ⍀ 10 30° A Figure E9.2 E9.3 Find the power supplied and the power absorbed by each element in Fig. E9.3. 4⍀ ANSWER: PC = 0 W; PL = 0 W; P4⍀ = 1.78 W; P2⍀ = 2.22 W; PCS = -4 W. 2⍀ –j2 ⍀ –j2 ⍀ j2 ⍀ 2 0° A Figure E9.3 [hint] Superposition is not applicable to power. Why? EXAMPLE 9.4 SOLUTION When determining average power, if more than one source is present in a network, we can use any of our network analysis techniques to find the necessary voltage and/or current to compute the power. However, we must remember that in general we cannot apply superposition to power. Consider the network shown in Fig. 9.5. We wish to determine the total average power absorbed and supplied by each element. From the figure we note that I2 = 12 / 30° 2 = 6 / 30° A and I3 = 12 / 30° - 6 / 0° j1 = 4.39 + j6 = 7.44 / -36.21° A j1 The power absorbed by the 2-⍀ resistor is P2 = 1 1 VM IM = (12)(6) = 36 W 2 2 irwin09_435-490hr.qxd 28-07-2010 12:00 Page 441 SECTION 9.2 I1 ± – 441 Figure 9.5 j1 ⍀ I3 Example RL circuit with two sources. I2 12 30° V AV E R AG E P OW E R ± – 2⍀ 6 0° V [hint] Under the following condition: I According to the direction of I3, the 6 / 0°-V source is absorbing power. The power it absorbs is given by 1 P6/ 0° = VM IM cos(v - i) 2 1 = (6)(7.44) cos [0° - (-36.21°)] 2 = 18 W ± V – if P ⫽ IV is positive, power is being absorbed. If P ⫽ IV is negative, power is being generated. At this point an obvious question arises: how do we know whether the 6 / 0°-V source is supplying power to the remainder of the network or absorbing it? The answer to this question is actually straightforward. If we employ our passive sign convention that was adopted in the earlier chapters—that is, if the current reference direction enters the positive terminal of the source and the answer is positive—the source is absorbing power. If the answer is negative, the source is supplying power to the remainder of the circuit. A generator sign convention could have been used, and under this condition the interpretation of the sign of the answer would be reversed. Note that once the sign convention is adopted and used, the sign for average power will be negative only if the angle difference is greater than 90° Ai.e., @ v - i @ 7 90°B. To obtain the power supplied to the network, we compute I1 as I1 = I2 + I3 = 6 / 30° + 7.44 / -36.21° = 11.29 / -7.10° A Therefore, the power supplied by the 12 / 30°-V source using the generator sign convention is 1 (12)(11.29) cos (30° + 7.10°) 2 = 54 W PS = and hence the power absorbed is equal to the power supplied. Learning Assessments E9.4 Determine the total average power absorbed and supplied by each element in the network in Fig. E9.4. –j2 ⍀ 12 0° A Figure E9.4 4⍀ ± – 4 30° V ANSWER: PCS = -69.4 W; PVS = 19.8 W; P4⍀ = 49.6 W; PC = 0. irwin09_435-490hr.qxd 442 28-07-2010 CHAPTER 9 12:00 Page 442 S T E A D Y- S TAT E P O W E R A N A LY S I S E9.5 Given the network in Fig. E9.5, determine the total average power absorbed or supplied by each element. 2⍀ j2 ⍀ 24 0° V ± – ANSWER: P24/0° = -55.4 W; P12/0° = 5.5 W; P2⍀ = 22.2 W; P4⍀ = 27.7 W; PL = 0. 4⍀ ± – 12 0° V Figure E9.5 E9.6 Determine the average power absorbed by the 4-⍀ and 3-⍀ resistors in Fig. E9.6. 3⍀ j2 ⍀ ANSWER: P4⍀ = 9.86 W; P3⍀ = 0.91 W. 4⍀ j3 ⍀ + - –j2 ⍀ 12 0° V + - 3 10° A 6 0° V Figure E9.6 In our study of resistive networks, we addressed the problem of maximum power transfer to a resistive load. We showed that if the network excluding the load was represented by a Thévenin equivalent circuit, maximum power transfer would result if the value of the load resistor was equal to the Thévenin equivalent resistance Ai.e., RL = RTh B. We will now reexamine this issue within the present context to determine the load impedance for the network shown in Fig. 9.6 that will result in maximum average power being absorbed by the load impedance ZL . The equation for average power at the load is 9.3 Maximum Average Power Transfer ZTh Voc ± – IL PL = + VL ZL 1 V I cos AvL - iL B 2 L L The phasor current and voltage at the load are given by the expressions - IL = Voc Z Th + Z L 9.12 VL = Voc Z L Z Th + Z L 9.13 ac circuit Figure 9.6 Circuit used to examine maximum average power transfer. 9.11 where Z Th = RTh + jXTh 9.14 Z L = RL + jXL 9.15 and irwin09_435-490hr.qxd 28-07-2010 12:00 Page 443 SECTION 9.3 M A X I M U M AV E R AG E P OW E R T R A N S F E R [hint] The magnitude of the phasor current and voltage are given by the expressions IL = VL = Voc 9.16 C ARTh + RL B + AXTh + XL B D 1兾2 2 2 Voc AR2L + X2L B 1兾2 9.17 C ARTh + RL B + AXTh + XL B D 1兾2 2 2 The phase angles for the phasor current and voltage are contained in the quantity AvL - iL B. Note also that vL - iL = ZL and, in addition, cos ZL = AR2L RL + X2L B 9.18 1兾2 Substituting Eqs. (9.16) to (9.18) into Eq. (9.11) yields PL = V2oc RL 1 2 ARTh + RL B 2 + AXTh + XL B 2 9.19 which could, of course, be obtained directly from Eq. (9.16) using PL = 21 I 2L RL . Once again, a little forethought will save us some work. From the standpoint of maximizing PL , Voc is a constant. The quantity AXTh + XL B absorbs no power, and therefore any nonzero value of this quantity only serves to reduce PL . Hence, we can eliminate this term by selecting XL = -XTh . Our problem then reduces to maximizing PL = 443 V2oc RL 1 2 ARL + RTh B 2 This impedance-matching concept is an important issue in the design of high-speed computer chips and motherboards. For today’s high-speed chips with internal clocks running at about 3 GHz and motherboards with a bus speed above 1 GHz, impedance matching is necessary in order to obtain the required speed for signal propagation. Although this high-speed transmission line is based on a distributed circuit (discussed later in electrical engineering courses), the impedancematching technique for the transmission line is the same as that of the lumped parameter circuit for maximum average power transfer. 9.20 However, this is the same quantity we maximized in the purely resistive case by selecting RL = RTh . Therefore, for maximum average power transfer to the load shown in Fig. 9.6, ZL should be chosen so that Z L = RL + jXL = RTh - jXTh = Z *Th 9.21 Finally, if the load impedance is purely resistive Ai.e., XL = 0B, the condition for maximum average power transfer can be derived via the expression dPL = 0 dRL where PL is the expression in Eq. (9.19) with XL=0. The value of RL that maximizes PL under the condition XL=0 is RL = 2R2Th + X2Th 9.22 Problem-Solving Strategy Step 1. Remove the load ZL and find the Thévenin equivalent for the remainder of the circuit. Step 2. Construct the circuit shown in Fig. 9.6. Step 3. Select Z L = Z *Th = RTh - jXTh , and then I L = Voc兾2 RTh and the maximum average power transfer = 1 2 I L RTh = V2oc兾8 RTh . 2 Maximum Average Power Transfer irwin09_435-490hr.qxd 444 28-07-2010 CHAPTER 9 12:00 Page 444 S T E A D Y- S TAT E P O W E R A N A LY S I S Given the circuit in Fig. 9.7a, we wish to find the value of ZL for maximum average power transfer. In addition, we wish to find the value of the maximum average power delivered to the load. EXAMPLE 9.5 SOLUTION To solve the problem, we form a Thévenin equivalent at the load. The circuit in Fig. 9.7b is used to compute the open-circuit voltage [hint] 4 / 0° (2) Voc = In this Thévenin analysis, 1. Remove ZL and find the voltage across the open terminals, Voc . 2. Determine the impedance ZTh at the open terminals with all independent sources made zero. 3. Construct the following circuit and determine I and PL. 6 + j1 (4) = 5.26 / -9.46° V The Thévenin equivalent impedance can be derived from the circuit in Fig. 9.7c. As shown in the figure, 4(2 + j1) Z Th = = 1.41 + j0.43 ⍀ 6 + j1 Therefore, ZL for maximum average power transfer is Z L = 1.41 - j0.43 ⍀ With ZL as given previously, the current in the load is I ZTh I = ± Voc – Z*Th 5.26 / -9.46° 2.82 = 1.87 / -9.46° A Therefore, the maximum average power transferred to the load is PL = 1 2 1 I M RL = (1.87)2(1.41) = 2.47 W 2 2 j1 ⍀ j1 ⍀ + 4 0° A 2⍀ ZL 4⍀ 2⍀ 4 0° A 4⍀ Voc - (a) (b) j1 ⍀ Figure 9.7 Circuits for illustrating maximum average power transfer. EXAMPLE 9.6 SOLUTION 2⍀ 4⍀ ZTh (c) For the circuit shown in Fig. 9.8a, we wish to find the value of ZL for maximum average power transfer. In addition, let us determine the value of the maximum average power delivered to the load. We will first reduce the circuit, with the exception of the load, to a Thévenin equivalent circuit. The open-circuit voltage can be computed from Fig. 9.8b. The equations for the circuit are Vxœ + 4 = (2 + j4)I 1 Vxœ = -2I 1 irwin09_435-490hr.qxd 28-07-2010 12:00 Page 445 SECTION 9.3 M A X I M U M AV E R AG E P OW E R T R A N S F E R Solving for I1 , we obtain I1 = [hint] 1 / -45° When there is a dependent source, both Voc and Isc must be found and ZTh computed from the equation 12 The open-circuit voltage is then Voc = 2I 1 - 4 / 0° = 22 / -45° - 4 / 0° ZTh = Voc Isc = -3 - j1 = +3.16 / -161.57° V The short-circuit current can be derived from Fig. 9.8c. The equations for this circuit are Vxfl + 4 = (2 + j4)I - 2I sc -4 = -2I + (2 - j2)I sc Vxfl = -2AI - I sc B Solving these equations for Isc yields I sc = -(1 + j2) A The Thévenin equivalent impedance is then Z Th = Voc 3 + j1 = = 1 - j1 ⍀ I sc 1 + j2 Therefore, for maximum average power transfer the load impedance should be Z L = 1 + j1 ⍀ The current in this load ZL is then Voc -3 - j1 = = 1.58 / -161.57° A Z Th + Z L 2 IL = Hence, the maximum average power transferred to the load is 1 (1.58)2(1) 2 = 1.25 W PL = –j2 ⍀ j4 ⍀ - Vx - V"x 2⍀ + ± – ZL – ± V'x 2⍀ + ± – 4 0° V –j2 ⍀ j4 ⍀ + V'x 2⍀ + Vx –j2 ⍀ I1 j4 ⍀ - Voc – ± V"x ± – I Isc – ± 4 0° V - (a) Figure 9.8 Circuits for illustrating maximum average power transfer. 445 (b) (c) 4 0° V irwin09_435-490hr.qxd 446 28-07-2010 CHAPTER 9 12:00 Page 446 S T E A D Y- S TAT E P O W E R A N A LY S I S Learning Assessments E9.7 Given the network in Fig. E9.7, find ZL for maximum average power transfer and the maximum average power transferred to the load. ANSWER: ZL = 1 + j1 ⍀; PL = 45 W. 24 0° V –± – j2 ⍀ – ± 12 0° V 2⍀ ZL j2 ⍀ Figure E9.7 E9.8 Find ZL for maximum average power transfer and the maximum average power transferred to the load in the network in Fig. E9.8. 24 0° V ANSWER: ZL = 2 - j2 ⍀; PL = 45 W. 12 0° V + + –j2 ⍀ 2⍀ j2 ⍀ ZL Figure E9.8 E9.9 Determine ZL for maximum average power transfer and the value of the maximum average power transferred to ZL in Fig. E9.9. 3⍀ j2 ⍀ 4⍀ ANSWER: ZL = 4.79 - j1.68⍀; PL = 14.26 W. j3 ⍀ + - ZL 12 0° V + - 3 10° A 6 0° V Figure E9.9 E9.10 Find ZL for maximum average power transfer and the value of the maximum average power transferred to ZL in Fig. E9.10. 3⍀ j2 ⍀ 4⍀ Ix j3 ⍀ + – Figure E9.1o 2 Ix ZL + - 6 0° V 3 10° A ANSWER: ZL = 5.67 - j 2.2 ⍀; PL = 9.29 W. irwin09_435-490hr.qxd 28-07-2010 12:00 Page 447 SECTION 9.4 E F F E C T I V E O R R M S VA L U E S In the preceding sections of this chapter, we have shown that the average power absorbed by a resistive load is directly dependent on the type, or types, of sources that are delivering power to the load. For example, if the source was dc, the average power absorbed was I2R, and if the source was sinusoidal, the average power was 1兾2 I 2M R. Although these two types of waveforms are extremely important, they are by no means the only waveforms we will encounter in circuit analysis. Therefore, a technique by which we can compare the effectiveness of different sources in delivering power to a resistive load would be quite useful. To accomplish this comparison, we define what is called the effective value of a periodic waveform, representing either voltage or current. Although either quantity could be used, we will employ current in the definition. Hence, we define the effective value of a periodic current as a constant or dc value, which as current would deliver the same average power to a resistor R. Let us call the constant current Ieff . Then the average power delivered to a resistor as a result of this current is 447 9.4 Effective or rms Values P = I 2eff R Similarly, the average power delivered to a resistor by a periodic current i(t) is P = t0 + T 1 i2(t)R dt T 3t0 Equating these two expressions, we find that Ieff = t0 + T 1 i2(t) dt C T 3t0 9.23 Note that this effective value is found by first determining the square of the current, then computing the average or mean value, and finally taking the square root. Thus, in “reading” the mathematical Eq. (9.23), we are determining the root mean square, which we abbreviate as rms, and therefore Ieff is called Irms . Since dc is a constant, the rms value of dc is simply the constant value. Let us now determine the rms value of other waveforms. The most important waveform is the sinusoid, and therefore, we address this particular one in the following example. We wish to compute the rms value of the waveform i(t) = IM cos (t-), which has a period of T=2/. Substituting these expressions into Eq. (9.23) yields Irms = c Using the trigonometric identity 1 1 + cos 2 2 2 we find that the preceding equation can be expressed as Irms = IM e 2兾 1兾2 1 1 c + cos (2t - 2) d dt f 2 30 2 2 Since we know that the average or mean value of a cosine wave is zero, Irms = IM a = IM c 2兾 1兾2 1 dt b 2 30 2 IM t 2兾 1兾2 a b2 d = 2 2 0 12 9.7 SOLUTION T 1兾2 1 I 2M cos2(t - ) dt d T 30 cos2 = EXAMPLE 9.24 irwin09_435-490hr.qxd 448 28-07-2010 CHAPTER 9 12:00 Page 448 S T E A D Y- S TAT E P O W E R A N A LY S I S Therefore, the rms value of a sinusoid is equal to the maximum value divided by 12 . Hence, a sinusoidal current with a maximum value of IM delivers the same average power to a resistor R as a dc current with a value of IM兾12 . Recall that earlier a phasor X was defined as XM / for a sinusoidal wave of the form XM cos (t + ). This phasor can also be represented as XM兾12 / if the units are given in rms. For example, 120 / 30° V rms is equivalent to 170 / 30° V. On using the rms values for voltage and current, the average power can be written, in general, as P = Vrms Irms cos Av - i B 9.25 The power absorbed by a resistor R is P = I 2rms R = V2rms R 9.26 In dealing with voltages and currents in numerous electrical applications, it is important to know whether the values quoted are maximum, average, rms, or what. We are familiar with the 120-V ac electrical outlets in our home. In this case, the 120 V is the rms value of the voltage in our home. The maximum or peak value of this voltage is 120 12 = 170 V. The voltage at our electrical outlets could be written as 170 cos 377t V. The maximum or peak value must be given if we write the voltage in this form. There should be no question in our minds that this is the peak value. It is common practice to specify the voltage rating of ac electrical devices in terms of the rms voltage. For example, if you examine an incandescent light bulb, you will see a voltage rating of 120 V, which is the rms value. For now we will add an rms to our voltages and currents to indicate that we are using rms values in our calculations. EXAMPLE 9.8 SOLUTION Determine the rms value of the current waveform in Fig. 9.9 and use this value to compute the average power delivered to a 2-⍀ resistor through which this current is flowing. The current waveform is periodic with a period of T=4 s. The rms value is 2 4 1兾2 1 Irms = e c (4)2 dt + (-4)2 dt d f 4 30 32 = c 2 4 1兾2 1 a 16t 2 + 16t 2 b d 4 0 2 = 4A The average power delivered to a 2-⍀ resistor with this current is P = I 2rms R = (4)2(2) = 32 W Figure 9.9 Waveform used to illustrate rms values. Current (A) i(t) 4 –2 0 –4 2 4 6 t(s) irwin09_435-490hr.qxd 28-07-2010 12:00 Page 449 SECTION 9.4 E F F E C T I V E O R R M S VA L U E S EXAMPLE We wish to compute the rms value of the voltage waveform shown in Fig. 9.10. The waveform is periodic with period T = 3 s. The equation for the voltage in the time frame 0 ⱕ t ⱕ 3 s is SOLUTION 9.9 0 6 tⱕ1s 1 6 tⱕ2s 2 6 tⱕ3s 4t V v(t) = c 0 V -4t + 8 V The rms value is Vrms = e = c 1 2 3 1兾2 1 c (4t)2 dt + (0)2 dt + (8 - 4t)2 dtd f 3 30 31 32 1 16t3 1 64t2 16t3 3 1兾2 2 + a 64t a + b2 bd 3 3 0 2 3 2 = 1.89 V v(t) (V) Figure 9.10 Waveform used to illustrate rms values. 4 0 1 2 3 4 5 t(s) –4 Learning Assessments E9.11 The current waveform in Fig. E9.11 is flowing through a 4- Æ resistor. Compute the average power delivered to the resistor. ANSWER: P = 32 W. i(t) (A) 4 2 0 Figure E9.11 2 4 6 8 10 12 t(s) E9.12 The current waveform in Fig. E9.12 is flowing through a 10-⍀ resistor. Determine the average power delivered to the resistor. i(t) (A) 4 4 0 Figure E9.12 –4 2 6 12 8 10 449 t(s) ANSWER: P = 80 W. irwin09_435-490hr.qxd 450 28-07-2010 CHAPTER 9 12:00 Page 450 S T E A D Y- S TAT E P O W E R A N A LY S I S E9.13 The voltage across a 2-W resistor is given by the waveform in Fig. E9.13. Find the average power absorbed by the resistor. ANSWER: P = 38.22 W. v(t) (V) 10 2 4 6 Figure E9.13 8 10 12 14 16 18 20 t(s) –10 E9.14 Compute the rms value of the voltage waveform shown in Fig. E9.14. v(t) (V) ANSWER: Vrms = 1.633 V. 4 0 Figure E9.14 9.5 The Power Factor 2 4 6 8 t(s) The power factor is a very important quantity. Its importance stems in part from the economic impact it has on industrial users of large amounts of power. In this section we carefully define this term and then illustrate its significance via some practical examples. In Section 9.4 we showed that a load operating in the ac steady state is delivered an average power of P = Vrms Irms cos Av - i B We will now further define the terms in this important equation. The product Vrms Irms is referred to as the apparent power. Although the term cos Av - i B is a dimensionless quantity, and the units of P are watts, apparent power is normally stated in volt-amperes (VA) or kilovolt-amperes (kVA) to distinguish it from average power. We now define the power factor (pf) as the ratio of the average power to the apparent power; that is, pf = P = cos Av - i B Vrms Irms 9.27 where cos Av - i B = cos ZL 9.28 The angle v - i = ZL is the phase angle of the load impedance and is often referred to as the power factor angle. The two extreme positions for this angle correspond to a purely resistive load where ZL = 0 and the pf is 1, and the purely reactive load where ZL = ; 90° and the pf is 0. It is, of course, possible to have a unity pf for a load containing R, L, and C elements if the values of the circuit elements are such that a zero phase angle is obtained at the particular operating frequency. There is, of course, a whole range of power factor angles between ;90° and 0°. If the load is an equivalent RC combination, then the pf angle lies between the limits -90° 6 ZL 6 0°. On the other hand, if the load is an equivalent RL combination, then the pf angle lies between the limits 0 6 ZL 6 90°. Obviously, confusion in identifying the type of load could result, due to the fact that cos ZL = cos A-ZL B. To circumvent this problem, the pf is said to be either leading or lagging, where these two terms refer to the phase of the current with respect to the voltage. Since the current leads the voltage in an RC load, the load has a leading pf. irwin09_435-490hr.qxd 28-07-2010 12:00 Page 451 SECTION 9.5 T H E P OW E R FAC TO R 451 In a similar manner, an RL load has a lagging pf; therefore, load impedances of ZL = 1 - j1 ⍀ and ZL = 2 + j1 ⍀ have power factors of cos(-45°) = 0.707 leading and cos(26.57°) = 0.894 lagging, respectively. An industrial load consumes 88 kW at a pf of 0.707 lagging from a 480-V rms line. The transmission line resistance from the power company’s transformer to the plant is 0.08 ⍀. Let us determine the power that must be supplied by the power company (a) under present conditions and (b) if the pf is somehow changed to 0.90 lagging. (It is economically advantageous to have a power factor as close to one as possible.) a. The equivalent circuit for these conditions is shown in Fig. 9.11. Using Eq. (9.27), we EXAMPLE 9.10 SOLUTION obtain the magnitude of the rms current into the plant: Irms = = PL (pf)AVrms B [hint] (88)A103 B Technique 1. Given PL, pf, and Vrms, determine Irms. 2. Then PS = PL + I2rms Rline, where Rline is the line resistance. (0.707)(480) = 259.3 A rms The power that must be supplied by the power company is PS = PL + (0.08)I 2rms = 88,000 + (0.08)(259.3)2 = 93.38 kW b. Suppose now that the pf is somehow changed to 0.90 lagging but the voltage remains constant at 480 V. The rms load current for this condition is Irms = = PL (pf)AVrms B (88)A103 B (0.90)(480) = 203.7 A rms Under these conditions, the power company must generate PS = PL + (0.08)I 2rms = 88,000 + (0.08)(203.7)2 = 91.32 kW Note carefully the difference between the two cases. A simple change in the pf of the load from 0.707 lagging to 0.90 lagging has had an interesting effect. Note that in the first case the power company must generate 93.38 kW in order to supply the plant with 88 kW of power because the low power factor means that the line losses will be high—5.38 kW. However, in the second case the power company need only generate 91.32 kW in order to supply the plant with its required power, and the corresponding line losses are only 3.32 kW. Figure 9.11 0.08 ⍀ Irms VS ± – 480 V rms PL=88 kW pf=0.707 lagging Example circuit for examining changes in power factor. irwin09_435-490hr.qxd 452 28-07-2010 CHAPTER 9 12:00 Page 452 S T E A D Y- S TAT E P O W E R A N A LY S I S Example 9.10 clearly indicates the economic impact of the load’s power factor. The cost of producing electricity for a large electric utility can easily be in the billions of dollars. A low power factor at the load means that the utility generators must be capable of carrying more current at constant voltage, and they must also supply power for higher I 2rms Rline losses than would be required if the load’s power factor were high. Since line losses represent energy expended in heat and benefit no one, the utility will insist that a plant maintain a high pf, typically 0.9 lagging, and adjust the rate it charges a customer that does not conform to this requirement. We will demonstrate a simple and economical technique for achieving this power factor correction in a future section. Learning Assessment E9.15 An industrial load consumes 100 kW at 0.707 pf lagging. The 60-Hz line voltage at the load is 480 / 0° V rms. The transmission-line resistance between the power company trans- ANSWER: Power saved is 3.771 kW. former and the load is 0.1 Æ. Determine the power savings that could be obtained if the pf is changed to 0.94 lagging. 9.6 In our study of ac steady-state power, it is convenient to introduce another quantity, which is commonly called complex power. To develop the relationship between this quantity and others we have presented in the preceding sections, consider the circuit shown in Fig. 9.12. The complex power is defined to be Complex Power S = Vrms I *rms where I *rms refers to the complex conjugate of Irms ; that is, if I rms = Irms / i = IR + jII , then I *rms = Irms / -i = IR - jII . Complex power is then Irms + Vrms 9.29 Z S = Vrms / v Irms / -i = Vrms Irms / v - i z - Figure 9.12 Circuit used to explain power relationships. or 9.30 S = Vrms Irms cos Av - i B + jVrms Irms sin Av - i B 9.31 where, of course, v - i = Z . We note from Eq. (9.31) that the real part of the complex power is simply the real or average power. The imaginary part of S we call the reactive or quadrature power. Therefore, complex power can be expressed in the form S = P + jQ where 9.32 P = Re(S) = Vrms Irms cos Av - i B 9.33 Q = Im(S) = Vrms Irms sin Av - i B 9.34 As shown in Eq. (9.31), the magnitude of the complex power is what we have called the apparent power, and the phase angle for complex power is simply the power factor angle. Complex power, like apparent power, is measured in volt-amperes, real power is measured in watts, and to distinguish Q from the other quantities, which in fact have the same dimensions, it is measured in volt-amperes reactive, or var. Now let’s examine the expressions in Eqs. (9.33) and (9.34) in more detail for our three basic circuit elements: R, L, and C. For a resistor, v - i = 0°, cos Av - i B = 1, and irwin09_435-490hr.qxd 28-07-2010 12:00 Page 453 SECTION 9.6 COMPLEX POWER sin Av - i B = 0. As a result, a resistor absorbs real power (P 7 0) but does not absorb any reactive power (Q = 0). For an inductor, v - i = 90° and P = Vrms Irms cos (90°) = 0 Q = Vrms Irms sin (90°) 7 0 An inductor absorbs reactive power but does not absorb real power. Repeating for a capacitor, we get v - i = -90° and P = Vrms Irms cos (-90°) = 0 Q = Vrms Irms sin (-90°) 6 0 A capacitor does not absorb any real power; however, the reactive power is now negative. How do we interpret the negative reactive power? Refer to Fig. 9.12 and note that the voltage and current are specified such that they satisfy the passive sign convention. In this case, the product of the voltage and current gives us the power absorbed by the impedance in that figure. If the reactive power absorbed by the capacitor is negative, then the capacitor must be supplying reactive power. The fact that capacitors are a source of reactive power will be utilized in the next section, on power factor correction. We see that resistors absorb only real power, while inductors and capacitors absorb only reactive power. What is a fundamental difference between these elements? Resistors only absorb energy. On the other hand, capacitors and inductors store energy and then release it back to the circuit. Since inductors and capacitors absorb only reactive power and not real power, we can conclude that reactive power is related to energy storage in these elements. Now let’s substitute Vrms = I rms * Z into Eq. (9.29). Multiplying I rms * I *rms = Irms / i * Irms / -i yields I 2rms . The complex power absorbed by an impedance can be obtained by multiplying the square of the rms magnitude of the current flowing through that impedance by the impedance. S = Vrms I *rms = AI rms ZBI *rms = I rms I *rms Z = I 2rms Z = I 2rms(R + jX) = P + jQ 9.35 Instead of substituting for Vrms in Eq. (9.29), let’s substitute for I rms : S = Vrms I *rms = Vrms a Vrms * V2rms b = = V2rms Y* = V2rms(G + jB)* = P + jQ Z Z* 9.36 This expression tells us that we can calculate the complex power absorbed by an admittance by multiplying the square of the rms magnitude of the voltage across the admittance by the conjugate of the admittance. Suppose the box in Fig. 9.12 contains a capacitor. The admittance for a capacitor is jC. Plugging into the equation above yields S = V2rms(jC)* = -jCV 2rms 9.37 Note the negative sign on the complex power. This agrees with our previous statement that a capacitor does not absorb real power but is a source of reactive power. The diagrams in Fig. 9.13 further explain the relationships among the various quantities of power. As shown in Fig. 9.13a, the phasor current can be split into two components: one Im Im Vrms s s co – ( v I rm v i) Irms sin (v–i) v–i i Im S Irms s ±Q v–i P v–i Re –Q S S= Z 2 m I r Q=I2rmsIm (Z) v–i P= I2rmsRe (Z) Re (a) Figure 9.13 Diagram for illustrating power relationships. (b) (c) Re 453 irwin09_435-490hr.qxd 454 28-07-2010 CHAPTER 9 12:00 Page 454 S T E A D Y- S TAT E P O W E R A N A LY S I S that is in phase with Vrms and one that is 90° out of phase with Vrms . Eqs. (9.33) and (9.34) illustrate that the in-phase component produces the real power, and the 90° component, called the quadrature component, produces the reactive or quadrature power. In addition, Eqs. (9.33) and (9.34) indicate that tan Av - i B = Q P 9.38 which relates the pf angle to P and Q in what is called the power triangle. The relationships among S, P, and Q can be expressed via the diagrams shown in Figs. 9.13b and c. In Fig. 9.13b we note the following conditions. If Q is positive, the load is inductive, the power factor is lagging, and the complex number S lies in the first quadrant. If Q is negative, the load is capacitive, the power factor is leading, and the complex number S lies in the fourth quadrant. If Q is zero, the load is resistive, the power factor is unity, and the complex number S lies along the positive real axis. Fig. 9.13c illustrates the relationships expressed by Eqs. (9.35) to (9.37) for an inductive load. In Chapter 1, we introduced Tellegen’s theorem, which states that the sum of the powers absorbed by all elements in an electrical network is zero. Based on this theorem, we can also state that complex power is conserved in an ac network—the total complex power delivered to any number of individual loads is equal to the sum of the complex powers delivered to the loads, regardless of how loads are interconnected. Problem-Solving Strategy Determining P or S If v(t) and i(t) are known and we wish to find P given an impedance Z / = R + jX, two viable approaches are as follows: Step 1. Determine V and I and then calculate P = Vrms Irms cos or P = Vrms Irms cos Av - i B Step 2. Use I to calculate the real part of S—that is, P = Re(S) = I 2R The latter method may be easier to calculate than the former. However, if the imaginary part of the impedance, X, is not zero, then P Z V2 R which is a common mistake. Furthermore, the P and Q portions of S are directly related to Z / and provide a convenient way in which to relate power, current, and impedance. That is, Q tan = P S = I2Z The following example illustrates the usefulness of S. EXAMPLE 9.11 SOLUTION A load operates at 20 kW, 0.8 pf lagging. The load voltage is 220 / 0° V rms at 60 Hz. The impedance of the line is 0.09 + j0.3 ⍀. We wish to determine the voltage and power factor at the input to the line. The circuit diagram for this problem is shown in Fig. 9.14. As illustrated in Fig. 9.13, 20,000 P P = = = 25,000 VA S = cos pf 0.8 irwin09_435-490hr.qxd 28-07-2010 12:00 Page 455 SECTION 9.6 COMPLEX POWER 455 Therefore, at the load SL = 25,000 / = 25,000 / 36.87° = 20,000 + j15,000 VA Since SL = VL I *L IL = c 25,000 / 36.87° * d 220 / 0° = 113.64 / -36.87° A rms The complex power losses in the line are Sline = I 2L Z line = (113.64)2(0.09 + j0.3) = 1162.26 + j3874.21 VA As stated earlier, complex power is conserved, and, therefore, the complex power at the generator is SS = SL + Sline = 21,162.26 + j18,874.21 = 28,356.25 / 41.73° VA Hence, the generator voltage is @ SS @ 28,356.25 = IL 113.64 = 249.53 V rms VS = and the generator power factor is cos (41.73°) = 0.75 lagging We could have solved this problem using KVL. For example, we calculated the load current as I L = 113.64 / -36.87° A rms Hence, the voltage drop in the transmission line is Vline = A113.64 / -36.87°B(0.09 + j0.3) = 35.59 / 36.43° V rms Therefore, the generator voltage is VS = 220 / 0° + 35.59 / 36.43° = 249.53 / 4.86° V rms Hence, the generator voltage is 249.53 V rms. In addition, v - i = 4.86° - (-36.87°) = 41.73° and therefore, [hint] 1. Use the given PL , cos , and VL rms to obtain SL and IL based on Eqs. (9.33) and (9.29), respectively. 2. Use IL and Zline to obtain Sline using Eq. (9.35). 3. Use SS = Sline + SL . 4. VS = SS兾I*L yields VS and v. Since VS = VS / v and i is the phase of IL , pf = cos Av - i B. pf = cos (41.73°) = 0.75 lagging 0.09 ⍀ j0.3 ⍀ Figure 9.14 Example circuit for power analysis. IL VS ± – 20 kW 0.8 pf lagging 220 0° V rms irwin09_435-490hr.qxd 456 28-07-2010 CHAPTER 9 12:00 Page 456 S T E A D Y- S TAT E P O W E R A N A LY S I S EXAMPLE Two networks A and B are connected by two conductors having a net impedance of Z = 0 + j1 ⍀, as shown in Fig. 9.15. The voltages at the terminals of the networks are VA = 120 / 30° V rms and VB = 120 / 0° V rms. We wish to determine the average power flow between the networks and identify which is the source and which is the load. 9.12 SOLUTION As shown in Fig. 9.15, VA - VB Z 120 / 30° - 120 / 0° I = = j1 = 62.12 / 15° A rms The power delivered by network A is PA = ∑VA∑ ∑I∑ cos AVA - I B = (120)(62.12) cos (30° - 15°) = 7200.4 W The power absorbed by network B is PB = ∑VB∑ ∑I∑ cos AVB - I B = (120)(62.12) cos (0° - 15°) = 7200.4 W If the power flow had actually been from network B to network A, the resultant signs on PA and PB would have been negative. Figure 9.15 I Network used in Example 9.12. + Network A Z + VA VB - - Network B Learning Assessments E9.16 An industrial load requires 40 kW at 0.84 pf lagging. The load voltage is 220 / 0° V rms at 60 Hz. The transmission-line impedance is 0.1+j0.25 ⍀. Determine the real and reactive power losses in the line and the real and reactive power required at the input to the transmission line. ANSWER: Pline = 4.685 kW; Qline = 11.713 kvar; PS = 44.685 kW; QS = 37.55 kvar. E9.17 A load requires 60 kW at 0.85 pf lagging. The 60-Hz line voltage at the load is 220 / 0° V rms. If the transmission-line impedance is 0.12+j0.18 ⍀, determine the line voltage and power factor at the input. ANSWER: Vin = 284.6 / 5.8° V rms; pfin = 0.792 lagging. E9.18 The source in Fig. E9.18 supplies 40 kW at a power factor of 0.9 lagging. The real and reactive losses of the transmission-line feeder are 1.6 kW and 2.1 kvar, respectively. Find the load voltage and the real and reactive power absorted by the load. ANSWER: VL = 416.83 ⬔ - 1.62⬚ V; PL = 38.4 kW; QL = 17.27 kvar. R jX + + – Figure E9.18 440 0° V rms VL - Load irwin09_435-490hr.qxd 28-07-2010 12:00 Page 457 SECTION 9.7 P OW E R FAC TO R C O R R E C T I O N E9.19 Find the power factor of the source and vS (t) in Fig. E9.19. if f = 60 Hz. ANSWER: pftn = 0.9457 lagging; vS (t) = 765.94 cos(377t - 7.77°) V. j0.4 ⍀ 0.2 ⍀ 50 kW 0.82 lagging ± v S – 50 kVA 0.95 leading 457 + 480 0° V rms - Figure E9.19 Industrial plants that require large amounts of power have a wide variety of loads. However, by nature the loads normally have a lagging power factor. In view of the results obtained in Example 9.10, we are naturally led to ask whether there is any convenient technique for raising the power factor of a load. Since a typical load may be a bank of induction motors or other expensive machinery, the technique for raising the pf should be an economical one to be feasible. To answer the question we pose, consider the diagram in Fig. 9.16. A typical industrial load with a lagging pf is supplied by an electrical source. Also shown is the power triangle for the load. The load pf is cos Aold B. If we want to improve the power factor, we need to reduce the angle shown on the power triangle in Fig. 9.16. From Eq. (9.38) we know that the tangent of this angle is equal to the ratio of Q to P. We could decrease the angle by increasing P. This is not an economically attractive solution because our increased power consumption would increase the monthly bill from the electric utility. QL=Qold IL Electrical source + VL - vL–iL=old PL=Pold The other option we have to reduce this angle is to decrease Q. How can we decrease Q? Recall from a previous section that a capacitor is a source of reactive power and does not absorb real power. Suppose we connect a capacitor in parallel with our industrial load as shown in Fig. 9.17. The corresponding power triangles for this diagram are also shown in Fig. 9.17. Let’s define Sold = Pold + jQold = ∑Sold∑ / old and Snew = Pold + jQnew = ∑Snew∑ / new Then with the addition of the capacitor, Snew = Sold + Scap Therefore, Scap = Snew - Sold = (Pnew - jQnew) - (Pold + jQold) = j (Qnew - Qold) = j (Qcap) Power Factor Correction Figure 9.16 Diagram for power factor correction. SL=Sold Industrial load with lagging pf 9.7 irwin09_435-490hr.qxd 458 28-07-2010 CHAPTER 9 12:00 Page 458 S T E A D Y- S TAT E P O W E R A N A LY S I S Recall from Eqs. (9.36) and (9.37) that in general S = V2rms/Z* and for a capacitor Z* = -1/jC so that Scap = Qcap = -jCV2rms This equation can be used to find the required value of C in order to achieve the new specified power factor defined by the new power factor angle illustrated in Fig. 9.17. Hence, we can obtain a particular power factor for the total load (industrial load and capacitor) simply by judiciously selecting a capacitor and placing it in parallel with the original load. In general, we want the power factor to be large, and therefore the power factor angle must be small [i.e., the larger the desired power factor, the smaller the angle AvL - iT B ]. IT Figure 9.17 Power factor correction diagram including capacitor. Electrical source IL + C Industrial load with lagging pf VL - QL=Qold SL=Sold Snew vL– iL=old Qnew new PL=Pold vL– iT=new PL=Pold Qcap EXAMPLE 9.13 Every month our electrical energy provider sends us a bill for the amount of electrical energy that we have consumed. The rate is often expressed in cents per kWh and consists of at least two components: (1) the demand charge, which covers the cost of lines, poles, transformers, and so on, and (2) the energy charge, which covers the cost to produce electric energy at power plants. The energy charge is the subject of the deregulation of the electric utility industry where you, as a customer, choose your energy provider. It is common for an industrial facility operating at a poor power factor to be charged more by the electric utility providing electrical service. Let’s suppose that our industrial facility operates at 277 V rms and requires 500 kW at a power factor of 0.75 lagging. Assume an energy charge of 2¢ per kWh and a demand charge of $3.50 per kW per month if the power factor is between 0.9 lagging and unity and $5 per kVA per month if the power factor is less than 0.9 lagging. The monthly energy charge is 500 * 24 * 30 * $0.02 = $7200. Let’s calculate the monthly demand charge with the 0.75 lagging power factor. The complex power absorbed by the industrial facility is Sold = 500 / cos-1 (0.75) = 666.67 / 41.4° = 500 + j441 kVA 0.75 irwin09_435-490hr.qxd 28-07-2010 12:00 Page 459 SECTION 9.7 P OW E R FAC TO R C O R R E C T I O N 459 The monthly demand charge is 666.67 * $5 = $3333.35. The total bill from the energy provider is $7200 + $3333.35 = $10533.35 per month. Let’s consider installing a capacitor bank, as shown in Fig. 9.18, to correct the power factor and reduce our demand charge. The demand charge is such that we only need to correct the power factor to 0.9 lagging. The monthly demand charge will be the same whether the power factor is corrected to 0.9 or unity. The complex power absorbed by the industrial facility and capacitor bank will be Snew = 500 / cos-1 (0.9) = 555.6 / 25.84° = 500 + j242.2 kVA 0.9 The monthly demand charge for our industrial facility with the capacitor bank is 500 * $3.50 = $1750 per month. The average power absorbed by our capacitor bank is negligible compared to the average power absorbed by the industrial facility, so our monthly energy charge remains $7200 per month. With the capacitor bank installed, the total bill from the energy provider is $7200 + $1750 = $8950 per month. How many kvars of capacitance do we need to correct the power factor to 0.9 lagging? Snew - Sold = Scap = (500 + j242.2) - (500 + j441) = -j198.8 kvar Let’s assume that it costs $100 per kvar to install the capacitor bank at the industrial facility for an installation cost of $19,880. How long will it take to recover the cost of installing the capacitor bank? The difference in the monthly demand charge without the bank and with the bank is $3333.35 - $1750 = $1583.35. Dividing this value into the cost of installing the bank yields $19,880兾$1583.35 = 12.56 months. Figure 9.18 A bank of capacitors. (Courtesy of Jeremy Nelms, Talquin Electric Cooperative, Inc.) Plastic kayaks are manufactured using a process called rotomolding, which is diagrammed in Fig. 9.19. Molten plastic is injected into a mold, which is then spun on the long axis of the kayak until the plastic cools, resulting in a hollow one-piece craft. Suppose that the induction motors used to spin the molds consume 50 kW at a pf of 0.8 lagging from a 220 / 0°-V rms, 60-Hz line. We wish to raise the pf to 0.95 lagging by placing a bank of capacitors in parallel with the load. EXAMPLE 9.14 Figure 9.19 Kayak mold Induction motor Rotomolding manufacturing process. irwin09_435-490hr.qxd 460 28-07-2010 CHAPTER 9 12:00 Page 460 S T E A D Y- S TAT E P O W E R A N A LY S I S SOLUTION The circuit diagram for this problem is shown in Fig. 9.20. PL = 50 kW and since cos-1 0.8 = 36.87°, old = 36.87°. Therefore, Hence, Qold = Pold tan old = (50)A103 B(0.75) = 37.5 kvar Sold = Pold + jQold = 50,000 + j37,500 and Scap = 0 + jQcap Since the required power factor is 0.95, new = cos-1 Apfnew B = cos-1 (0.95) = 18.19° Then Qnew = Pold tan new = 50,000 tan (18.19°) = 16,430 var Hence Qnew - Qold = Qcap = -CV2 rms 16,430 - 37,500 = -CV2 rms Solving the equation for C yields C = 21,070 (377)(220)2 = 1155 F By using a capacitor of this magnitude in parallel with the industrial load, we create, from the utility’s perspective, a load pf of 0.95 lagging. However, the parameters of the actual load remain unchanged. Under these conditions, the current supplied by the utility to the kayak manufacturer is less and therefore they can use smaller conductors for the same amount of power. Or, if the conductor size is fixed, the line losses will be less since these losses are a function of the square of the current. Figure 9.20 Example circuit for power factor correction. IT + 220 0° V rms IC IL 50 kW pf=0.8 lagging - Problem-Solving Strategy Power Factor Correction Step 1. Find Qold from PL and old , or the equivalent pfold . Step 2. Find new from the desired pfnew . Step 3. Determine Qnew=Pold tan new . Step 4. Qnew-Qold=Qcap=–CV2 rms. C irwin09_435-490hr.qxd 28-07-2010 12:00 Page 461 SECTION 9.8 SINGLE-PHASE THREE-WIRE CIRCUITS 461 Learning Assessments E9.20 Compute the value of the capacitor necessary to change the power factor in Learning Assessment E9.16 to 0.95 lagging. E9.21 Find the value of capacity to be connected in parallel with the load in Fig. E9.21 to make ANSWER: C = 773 F. ANSWER: C = 546.2 F. the source power factor 0.95 leading, f = 60 Hz. + – 50 kW 0.82 lagging 480 0° V rms Figure E9.21 The single-phase three-wire ac circuit shown in Fig. 9.21 is an important topic because it is the typical ac power network found in households. Note that the voltage sources are equal; that is, Van = Vnb = V. Thus, the magnitudes are equal and the phases are equal (single phase). The line-to-line voltage Vab = 2Van = 2Vnb = 2V. Within a household, lights and small appliances are connected from one line to neutral n, and large appliances such as hot water heaters and air conditioners are connected line to line. Lights operate at about 120 V rms and large appliances operate at approximately 240 V rms. Let us now attach two identical loads to the single-phase three-wire voltage system using perfect conductors as shown in Fig. 9.21b. From the figure we note that I aA = V ZL and I bB = - V ZL KCL at point N is I nN = - AI aA + I bB B = -a V V b ZL ZL = 0 Note that there is no current in the neutral wire, and therefore it could be removed without affecting the remainder of the system; that is, all the voltages and currents would be unchanged. One is naturally led to wonder just how far the simplicity exhibited by this system will extend. For example, what would happen if each line had a line impedance, if the neutral conductor had an impedance associated with it, and if there were a load tied from line to line? To explore these questions, consider the circuit in Fig. 9.21c. Although we could examine this circuit using many of the techniques we have employed in previous chapters, the symmetry of the network suggests that perhaps superposition may lead us to some conclusions without having to resort to a brute-force assault. Employing superposition, we consider the two circuits in Figs. 9.21d and e. The currents in Fig. 9.21d are labeled arbitrarily. Because of the symmetrical relationship between Figs. 9.21d and e, the 9.8 Single-Phase Three-Wire Circuits irwin09_435-490hr.qxd 462 28-07-2010 CHAPTER 9 12:00 Page 462 S T E A D Y- S TAT E P O W E R A N A LY S I S IaA Figure 9.21 Single-phase three-wire system. a a ± V – ± V – ± V – ± V – (a) ± V – ZL Zn N B ZL Zline (b) Zline Ia ZLL ± V – ZL IbB b b Zline ZL InN n n A (c) Ic Zline –If Ib ± V – ZL Zn Id ZL ± V – ZL Zline Zn ZLL Ie (d) –Id ZLL ZL Zline If – Ic –Ib –Ia –Ic (e) currents in Fig. 9.21e correspond directly to those in Fig. 9.21d. If we add the two phasor currents in each branch, we find that the neutral current is again zero. A neutral current of zero is a direct result of the symmetrical nature of the network. If either the line impedances Zline or the load impedances ZL are unequal, the neutral current will be nonzero. We will make direct use of these concepts when we study three-phase networks in Chapter 11. EXAMPLE 9.15 SOLUTION A three-wire single-phase household circuit is shown in Fig. 9.22a. Use of the lights, stereo, and range for a 24-hour period is demonstrated in Fig. 9.22b. Let us calculate the energy use over the 24 hours in kilowatt-hours. Assuming that this represents a typical day and that our utility rate is $0.08/kWh, let us also estimate the power bill for a 30-day month. Applying nodal analysis to Fig. 9.22a yields I aA = I L + I R I bB = -I S - I R I nN = I S - I L The current magnitudes for each load can be found from the corresponding power levels as follows: irwin09_435-490hr.qxd 28-07-2010 12:00 Page 463 SECTION 9.8 IaA a IR Lights 120 W InN n ± – 120 0° V rms A IL ± – 120 0° V rms SINGLE-PHASE THREE-WIRE CIRCUITS Range 7200 W N on Lights Range IS IbB on Stereo off Stereo 24 W b off 12 A.M. on off 2 4 B (a) 6 8 10 2 12 P.M. PL 120 = = 1 A rms Van 120 IS = PS 24 = = 0.2 A rms Vnb 120 IR = PR 7200 = = 30 A rms Vab 240 The energy used is simply the integral of the power delivered by the two sources over the 24-hour period. Since the voltage magnitudes are constants, we can express the energy delivered by the sources as Ean = Van 3 IaA dt Enb = Vnb 3 -IbB dt The integrals of IaA and IbB can be determined graphically from Fig. 9.22b. 12a.m. IaA dt = 4IR + 15IL = 135 12a.m. 312a.m. 6 8 10 12 (b) IL = 312a.m. 4 -IbB dt = 8IS + 4IR = 121.6 Therefore, the daily energy for each source and the total energy is Ean = 16.2 kWh Enb = 14.6 kWh Etotal = 30.8 kWh Over a 30-day month, a $0.08/kWh utility rate results in a power bill of Cost = A30.8)A30)A0.08) = $73.92 The energy consumption is typically measured by meters, of the form shown in Fig. 9.23, which are a familiar sight on the outside of our homes. Figure 9.22 Household three-wire network and appliance usage. 463 irwin09_435-490hr.qxd 464 28-07-2010 CHAPTER 9 12:00 Page 464 S T E A D Y- S TAT E P O W E R A N A LY S I S Figure 9.23 Electric meters used to measure home energy consumption. (Left, Comstock/Punchstock; right, Robert Llewellyn/ Workbook Stock/Jupiter Images) 9.9 Safety Considerations Although this book is concerned primarily with the theory of circuit analysis, we recognize that, by this point in their study, most students will have begun to relate the theory to the electrical devices and systems that they encounter in the world around them. Thus, it seems advisable to depart briefly from the theoretical and spend some time discussing the very practical and important subject of safety. Electrical safety is a very broad and diverse topic that would require several volumes for a comprehensive treatment. Instead, we will limit our discussion to a few introductory concepts and illustrate them with examples. It would be difficult to imagine that anyone in our society could have reached adolescence without having experienced some form of electrical shock. Whether that shock was from a harmless electrostatic discharge or from accidental contact with an energized electrical circuit, the response was probably the same—an immediate and involuntary muscular reaction. In either case, the cause of the reaction is current flowing through the body. The severity of the shock depends on several factors, the most important of which are the magnitude, the duration, and the pathway of the current through the body. The effect of electrical shock varies widely from person to person. Figure 9.24 shows the general reactions that occur as a result of 60-Hz ac current flow through the body from hand to hand, with the heart in the conduction pathway. Observe that there is an intermediate range of current, from about 0.1 to 0.2 A, which is most likely to be fatal. Current levels in this range are apt to produce ventricular fibrillation, a disruption of the orderly contractions of the heart muscle. Recovery of the heartbeat generally does not occur without immediate medical intervention. Current levels above that fatal range tend to cause the heart muscle to contract severely, and if the shock is removed soon enough, the heart may resume beating on its own. The voltage required to produce a given current depends on the quality of the contact to the body and the impedance of the body between the points of contact. The electrostatic voltage such as might be produced by sliding across a car seat on a dry winter day may be on the order of 20,000 to 40,000 V, and the current surge on touching the door handle, on the order of 40 A. However, the pathway for the current flow is mainly over the body surface, and its duration is for only a few microseconds. Although that shock could be disastrous for some electronic components, it causes nothing more than mild discomfort and aggravation to a human being. Electrical appliances found about the home typically require 120 or 240 V rms for operation. Although the voltage level is small compared with that of the electrostatic shock, the potential for harm to the individual and to property is much greater. Accidental contact is more apt to result in current flow either from hand to hand or from hand to foot—either of which will subject the heart to shock. Moreover, the relatively slowly changing (low frequency) 60-Hz current tends to penetrate more deeply into the body as opposed to remaining on the surface as a rapidly changing (high frequency) current would tend to do. In addition, the energy source has the capability of sustaining a current flow without depletion. Thus, subsequent discussion will concentrate primarily on hazards associated with the 60-Hz ac power system. irwin09_435-490hr.qxd 28-07-2010 12:00 Page 465 SECTION 9.9 S A F E T Y C O N S I D E R AT I O N S 10 A 465 Figure 9.24 Severe burns, not fatal unless vital organs are burned Effects of electrical shock. (From C. F. Dalziel and W. R. Lee, “Lethal Electric Currents,” IEEE Spectrum, February 1969, pp. 44–50; and C. F. Dalziel, “Electric Shock Hazard,” IEEE Spectrum, February 1972, pp. 41–50.) 1A Heartbeat stops, may restart if shock is removed before death Ventricular fibrillation usually fatal without intervention 100 mA Breathing stops Muscular paralysis, severe pain, difficulty breathing Let-go threshold 10 mA 1 mA Painful Threshold of sensation The single-phase three-wire system introduced earlier is commonly, though not exclusively, used for electrical power distribution in residences. Two important aspects of this or any system that relate to safety were not mentioned earlier: circuit fusing and grounding. Each branch circuit, regardless of the type of load it serves, is protected from excessive current flow by circuit breakers or fuses. Receptacle circuits are generally limited to 20 amps and lighting circuits to 15 amps. Clearly, these cannot protect persons from lethal shock. The primary purpose of these current-limiting devices is to protect equipment. The neutral conductor of the power system is connected to ground (earth) at a multitude of points throughout the system and, in particular, at the service entrance to the residence. The connection to earth may be by way of a driven ground rod or by contact to a cold water pipe of a buried metallic water system. The 120-V branch circuits radiating from the distribution panel (fuse box) generally consist of three conductors rather than only two, as was shown in Fig. 9.21. The third conductor is the ground wire, as shown in Fig. 9.25. The ground conductor may appear to be redundant, since it plays no role in the normal operation of a load that might be connected to the receptacle. Its role is illustrated by the following example. Figure 9.25 Circuit breaker 120 V rms ± – A household receptacle. Other receptacles Neutral Ground conductor irwin09_435-490hr.qxd 466 28-07-2010 CHAPTER 9 12:00 Page 466 S T E A D Y- S TAT E P O W E R A N A LY S I S EXAMPLE 9.16 SOLUTION Joe has a workshop in his basement where he uses a variety of power tools such as drills, saws, and sanders. The basement floor is concrete, and being below ground level, it is usually damp. Damp concrete is a relatively good conductor. Unknown to Joe, the insulation on a wire in his electric drill has been nicked, and the wire is in contact with (or shorted to) the metal case of the drill, as shown in Fig. 9.26. Is Joe in any danger when using the drill? Without the ground conductor connected to the metal case of the tool, Joe would receive a severe, perhaps fatal, shock when he attempted to use the drill. The voltage between his hand and his feet would be 120 V, and the current through his body would be limited by the resistance of his body and of the concrete floor. Typically, the circuit breakers would not operate. However, if the ground conductor is present and properly connected to the drill case, the case remains at ground potential, the 120-V source becomes shorted to ground, the circuit breaker operates, and Joe lives to drill another hole. Short Figure 9.26 Faulty circuit, when the case of the tool is not grounded through the power cord. 120 V ± – Neutral Ground Concrete floor It was mentioned earlier that the circuit breaker or fuse cannot provide effective protection against shock. There is, however, a special type of device called a ground-fault interrupter (GFI) that can provide protection for personnel. This device detects current flow outside the normal circuit. Consider the circuit of Fig. 9.26. In the normal safe operating condition, the current in the neutral conductor must be the same as that in the line conductor. If at any time the current in the line does not equal the current in the neutral, then a secondary path has somehow been established, creating an unsafe condition. This secondary path is called a fault. For example, the fault path in Fig. 9.26 is through Joe and the concrete floor. The GFI detects this fault and opens the circuit in response. Its principle of operation is illustrated by the following example. EXAMPLE Let us describe the operation of a GFI. 9.17 SOLUTION Consider the action of the magnetic circuit in Fig. 9.27. Under normal operating conditions, i1 and i2 are equal, and if the coils in the neutral and line conductors are identical, as we learned in basic physics, the magnetic flux in the core will be zero. Consequently, no voltage will be induced in the sensing coil. If a fault should occur at the load, current will flow in the ground conductor and perhaps in the earth; thus, i1 and i2 will no longer be equal, the magnetic flux will not be zero, and a voltage will be induced in the sensing coil. That voltage can be used to activate a circuit breaker. This is the essence of the GFI device. i1 Load Figure 9.27 Ground-fault interrupter circuit. V ± – Sensing coil Neutral Ground Magnetic core i2 irwin09_435-490hr.qxd 28-07-2010 12:00 Page 467 SECTION 9.9 S A F E T Y C O N S I D E R AT I O N S 467 Ground-fault interrupters are available in the form of circuit breakers and also as receptacles. They are now required in branch circuits that serve outlets in areas such as bathrooms, basements, garages, and outdoor sites. The devices will operate at ground-fault currents on the order of a few milliamperes. Unfortunately, the GFI is a relatively new device, and electrical code requirements are generally not retroactive. Thus few older residences have them. Requirements for the installation and maintenance of electrical systems are meticulously defined by various codes that have been established to provide protection of personnel and property. Installation, alteration, or repair of electrical devices and systems should be undertaken only by qualified persons. The subject matter that we study in circuit analysis does not provide that qualification. The following examples illustrate the potential hazards that can be encountered in a variety of everyday situations. We begin by revisiting a situation described in a previous example. Suppose that a man is working on the roof of a mobile home with a hand drill. It is early in the day, the man is barefoot, and dew covers the mobile home. The ground prong on the electrical plug of the drill has been removed. Will the man be shocked if the “hot” electrical line shorts to the case of the drill? To analyze this problem, we must construct a model that adequately represents the situation described. In his book Medical Instrumentation (Boston: Houghton Mifflin, 1978), John G. Webster suggests the following values for resistance of the human body: Rskin(dry) = 15 k⍀, Rskin(wet) = 150 ⍀, Rlimb(arm or leg) = 100 ⍀, and Rtrunk = 200 ⍀. The network model is shown in Fig. 9.28. Note that since the ground line is open-circuited, a closed path exists from the hot wire through the short, the human body, the mobile home, and the ground. For the conditions stated previously, we assume that the surface contact resistances Rsc1 and Rsc2 are 150 ⍀ each. The body resistance, Rbody , consisting of arm, trunk, and leg, is 400 ⍀. The mobile home resistance is assumed to be zero, and the ground resistance, Rgnd, from the mobile home ground to the actual source ground is assumed to be 1 ⍀. Therefore, the magnitude of the current through the body from hand to foot would be EXAMPLE 9.18 SOLUTION 120 Rsc1 + Rbody + Rsc2 + Rgnd 120 = 701 = 171 mA I body = A current of this magnitude can easily cause heart failure. Additional protection would be provided if the circuit breaker were a ground-fault interrupter. Circuit breaker IS “Hot” black wire Figure 9.28 Model for Example 9.18. Short Rsc1 ± – IR Ground Neutral white wire Drill case Green wire Rgnd Rbody Rsc2 irwin09_435-490hr.qxd 468 28-07-2010 CHAPTER 9 12:00 Page 468 S T E A D Y- S TAT E P O W E R A N A LY S I S EXAMPLE 9.19 SOLUTION Two boys are playing basketball in their backyard. To cool off, they decide to jump into their pool. The pool has a vinyl lining, so the water is electrically insulated from the earth. Unknown to the boys, there is a ground fault in one of the pool lights. One boy jumps in and while standing in the pool with water up to his chest, reaches up to pull in the other boy, who is holding onto a grounded hand rail, as shown in Fig. 9.29a. What is the impact of this action? The action in Fig. 9.29a is modeled as shown in Fig. 9.29b. Note that since a ground fault has occurred, there exists a current path through the two boys. Assuming that the fault, pool, and railing resistances are approximately zero, the magnitude of the current through the two boys would be I = 120 A3Rarm B + 3A3Rwet contact B + Rtrunk 120 950 = 126 mA = This current level would cause severe shock in both boys. The boy outside the pool could experience heart failure. Figure 9.29 Diagrams used in Example 9.19. Circuit breaker Hand rail ± – Light (a) Circuit breaker Rfault ± – Rwet Rpool Rarm contact Rarm Rwet contact Rlight Rarm Rtrunk Rwet contact Rrailing (b) irwin09_435-490hr.qxd 28-07-2010 12:00 Page 469 SECTION 9.9 S A F E T Y C O N S I D E R AT I O N S EXAMPLE A patient in a medical laboratory has a muscle stimulator attached to her left forearm. Her heart rate is being monitored by an EKG machine with two differential electrodes over the heart and the ground electrode attached to her right ankle. This activity is illustrated in Fig. 9.30a. The stimulator acts as a current source that drives 150 mA through the muscle from the active electrode to the passive electrode. If the laboratory technician mistakenly decides to connect the passive electrode of the stimulator to the ground electrode of the EKG system to achieve a common ground, is there any risk? When the passive electrode of the stimulator is connected to the ground electrode of the EKG system, the equivalent network in Fig. 9.30b illustrates the two paths for the stimulator current: one through half an arm and the other through half an arm and the body. Using current division, the body current is 469 9.20 SOLUTION (150)A10-3 B(50) I body = 50 + 50 + 200 + 100 = 19 mA Figure 9.30 Therefore, a dangerously high level of current will flow from the stimulator through the body to the EKG ground. Diagrams used in Example 9.20. 150 mA Muscle stimulator Rarm Rarm Ibody EKG monitor Rtrunk (a) (b) A cardiac care patient with a pacing electrode has ignored the hospital rules and is listening to a cheap stereo. The stereo has an amplified 60-Hz hum that is very annoying. The patient decides to dismantle the stereo partially in an attempt to eliminate the hum. In the process, while he is holding one of the speaker wires, the other touches the pacing electrode. What are the risks in this situation? Let us suppose that the patient’s skin is damp and that the 60-Hz voltage across the speaker wires is only 10 mV. Then the circuit model in this case would be as shown in Fig. 9.31. Figure 9.31 Rdamp skin I Rarm ± – 10 mV Rleg Rtrunk Relectrode=0 ⍀ Circuit model for Example 9.21. EXAMPLE 9.21 SOLUTION irwin09_435-490hr.qxd 470 28-07-2010 CHAPTER 9 12:00 Page 470 S T E A D Y- S TAT E P O W E R A N A LY S I S The current through the heart would be I = (10)A10-3 B 150 + 100 + 200 = 22.2 A It is known that 10 A delivered directly to the heart is potentially lethal. EXAMPLE 9.22 SOLUTION While maneuvering in a muddy area, a crane operator accidentally touched a high-voltage line with the boom of the crane, as illustrated in Fig. 9.32a. The line potential was 7200 V. The neutral conductor was grounded at the pole. When the crane operator realized what had happened, he jumped from the crane and ran in the direction of the pole, which was approximately 10 m away. He was electrocuted as he ran. Can we explain this very tragic accident? The conditions depicted in Fig. 9.32a can be modeled as shown in Fig. 9.32b. The crane was at 7200 V with respect to earth. Therefore, a gradient of 720 V/m existed along the earth between the crane and the power pole. This earth between the crane and the pole is modeled as a resistance. If the man’s stride was about 1 m, the difference in potential between his feet was approximately 720 V. A man standing in the same area with his feet together was unharmed. Neutral conductor Figure 9.32 Illustrations used in Example 9.22. Power line (a) 7200 V ±– Earth resistance Crane Power pole 10 m (b) irwin09_435-490hr.qxd 28-07-2010 12:00 Page 471 SECTION 9.9 S A F E T Y C O N S I D E R AT I O N S 471 The examples of this section have been provided in an attempt to illustrate some of the potential dangers that exist when working or playing around electric power. In the worst case, failure to prevent an electrical accident can result in death. However, even nonlethal electrical contacts can cause such things as burns or falls. Therefore, we must always be alert to ensure not only our own safety, but also that of others who work and play with us. The following guidelines will help minimize the chances of injury: 1. Avoid working on energized electrical systems. 2. Always assume that an electrical system is energized unless you can absolutely verify 3. 4. 5. 6. [hint] Safety guidelines. that it is not. Never make repairs or alterations that are not in compliance with the provisions of the prevailing code. Do not work on potentially hazardous electrical systems alone. If another person is “frozen” to an energized electrical circuit, deenergize the circuit, if possible. If that cannot be done, use nonconductive material such as dry wooden boards, sticks, belts, and articles of clothing to separate the body from the contact. Act quickly but take care to protect yourself. When handling long metallic equipment, such as ladders, antennas, and so on, outdoors, be continuously aware of overhead power lines and avoid any possibility of contact with them. Learning Assessment E9.22 A woman is driving her car in a violent rainstorm. While she is waiting at an intersection, a power line falls on her car and makes contact. The power line voltage is 7200 V. (a) Assuming that the resistance of the car is negligible, what is the potential current ANSWER: (a) I = 463 mA, extremely dangerous; (b) she should be safe. through her body if, while holding the door handle with a dry hand, she steps out onto the wet ground? (b) If she remained in the car, what would happen? Safety when working with electric power must always be a primary consideration. Regardless of how efficient or expedient an electrical network is for a particular application, it is worthless if it is also hazardous to human life. The safety device shown in Fig. 9.33, which is also used for troubleshooting, is a proximity-type sensor that will indicate whether a circuit is energized by simply touching the conductor on the outside of the insulation. This device is typically carried by all electricians and is helpful when working on electric circuits. In addition to the numerous deaths that occur each year due to electrical accidents, fire damage that results from improper use of electrical wiring and distribution equipment amounts to millions of dollars per year. To prevent loss of life and damage to property, very detailed procedures and specifications have been established for the construction and operation of electrical systems to ensure their safe operation. The National Electrical Code ANSI C1 (ANSI—American National Figure 9.33 A modern safety or troubleshooting device. (Courtesy of Fluke Corporation. Reproduced with permission.) irwin09_435-490hr.qxd 472 28-07-2010 CHAPTER 9 12:00 Page 472 S T E A D Y- S TAT E P O W E R A N A LY S I S Standards Institute) is the primary guide. There are other codes, however: for example, the National Electric Safety Code, ANSI C2, which deals with safety requirements for public utilities. Underwriters Laboratory (UL) tests all types of devices and systems to ensure that they are safe for use by the general public. We find the UL label on all types of electrical equipment that is used in the home, such as appliances and extension cords. Electric energy plays a central role in our lives. It is extremely important to our general health and well-being. However, if not properly used, it can be lethal. 9.10 The following application-oriented examples illustrate a practical use of the material studied in this chapter. Application Examples APPLICATION EXAMPLE 9.23 SOLUTION For safety reasons the National Electrical Code restricts the circuit-breaker rating in a 120-V household lighting branch circuit to no more than 20 A. Furthermore, the code also requires a 25% safety margin for continuous-lighting loads. Under these conditions, let us determine the number of 100-W lighting fixtures that can be placed in one branch circuit. The model for the branch circuit is shown in Fig. 9.34. The current drawn by each 100-W bulb is Ibulb = 100兾120 = 0.833 A rms Using the safety margin recommendation, the estimated current drawn by each bulb is 25% higher, or Ibulb = A1.25BA0.83B = 1.04 A rms Therefore, the maximum number of fixtures on one circuit breaker is n = 20兾1.04 = 19 fixtures 20-A breaker Figure 9.34 20-A branch circuit for household lighting. a 120 0° V rms ± – 100 W 100 W 100 W 100 W n APPLICATION EXAMPLE 9.24 SOLUTION An electric lawn mower requires 12 A rms at 120 V rms but will operate down to 110 V rms without damage. At 110 V rms, the current draw is 13.1 A rms, as shown in Fig. 9.35. Give the maximum length extension cord that can be used with a 120-V rms power source if the extension cord is made from 1. 16-gauge wire (4 m⍀/ft) 2. 14-gauge wire (2.5 m⍀/ft) The voltage drop across the extension cord is Vcord = (2)(13.1)Rcord = 10 V rms or Rcord = 0.382 ⍀ irwin09_435-490hr.qxd 28-07-2010 12:00 Page 473 SECTION 9.10 A P P L I C AT I O N E X A M P L E S 473 If /cord is the length of the extension cord, then for 16-gauge wire we find /cord = Rcord = 95.5 feet 0.004 and for 14-gauge wire /cord Vcord=5 V rms – ± Rcord = = 152.8 feet 0.0025 Rcord 120 V rms Figure 9.35 Circuit model in Example 9.24. ± – Mower ± 110 V rms Rcord 13.1 A rms – – ± Vcord=5 V rms While sitting in a house reading a book, we notice that every time the air conditioner comes on, the lights momentarily dim. Let us investigate this phenomenon using the single-phase three-wire circuit shown in Fig. 9.36a and some typical current requirements for a 10,000-Btu/h air conditioner, assuming a line resistance of 0.5 ⍀. APPLICATION EXAMPLE 9.25 The 60-W light bulb can be roughly modeled by its equivalent resistance: SOLUTION Pbulb = V2an Rbulb or Rbulb = 240 ⍀ When the air conditioner unit first turns on, the current requirement is 40 A, as shown in Fig. 9.36b. As the compressor motor comes up to speed, the current requirement drops quickly to a steady-state value of 10 A, as shown in Fig. 9.36c. We will compare the voltage across the light fixture, VAN , both at turn-on and in steady state. Using superposition, let us first find that portion of VAN caused by the voltage sources. The appropriate circuit is shown in Fig. 9.36d. Using voltage division, we find that VAN1 = VAN a Rbulb b Rbulb + 2RL or VAN1 = 119.50 V rms Figure 9.36e will yield the contribution to VAN caused by the 10-A steady-state current. Using current division to calculate the current through the light bulb, we find that or VAN2 = - e IAB a RL b f Rbulb Rbulb + 2RL VAN2 = -4.98 V rms Therefore, the steady-state value of VAN is VAN = VAN1 + VAN2 = 114.52 V rms At start-up, our expression for VAN2 can be used with IAB = 40 A, which yields VAN2 = -19.92 V rms. The resulting value for VAN is VAN = VAN1 + VAN2 = 119.50 - 19.92 = 99.58 V rms The voltage delivered to the light fixture at startup is 13% lower than the steady-state value, resulting in a momentary dimming of the lights. [hint] Technique 1. Find the resistance of the light bulb. 2. Use a large current source to represent the transient current of the air conditioner and a small current source to represent the steady-state current. 3. Find the voltage drop across the light bulb during both the transient and steady-state operations. irwin09_435-490hr.qxd 474 28-07-2010 CHAPTER 9 12:00 Page 474 S T E A D Y- S TAT E P O W E R A N A LY S I S RL 0.5 ⍀ A RL 0.5 ⍀ A a a Rbulb ± – 120 0° V rms IAB Lights 60 W RL 0.5 ⍀ n 10,000 air Btu conditioner N b RL 0.5 ⍀ N RL 0.5 ⍀ b B B (b) (a) RL 0.5 ⍀ A a RL 0.5 ⍀ A a Rbulb ± – 120 0° V rms N ± – 120 0° V rms b IAB 240 ⍀ B 240 ⍀ 240 ⍀ N n IAB RL 0.5 ⍀ N 10 A ± – 120 0° V rms RL 0.5 ⍀ Rbulb RL 0.5 ⍀ n 10 A RL 0.5 ⍀ A a Rbulb ± – 120 0° V rms RL 0.5 ⍀ n 40 A ± – 120 0° V rms RL 0.5 ⍀ IAB 240 ⍀ n ± – 120 0° V rms ± – 120 0° V rms b RL 0.5 ⍀ B (c) b RL 0.5 ⍀ (d) B (e) Figure 9.36 Networks used in Example 9.25. APPLICATION EXAMPLE 9.26 SOLUTION Most clothes dryers operate from a 240- V rms outlet and have several temperature settings, such as low, medium, and high. Let’s examine the manner in which the dryer creates heat for drying and the way in which it regulates the temperature. First we will consider the heat itself. A simple model for this situation is shown in Fig. 9.37a, where a resistive heating element is connected across the 240-V rms supply. For a particular dryer, the resistance of the element is roughly 11 ⍀. The current through this element is Ihe = 240 = 21.81 A rms 11 9.39 Since the element is resistive, the voltage and current are in phase. The power dissipation is Phe = I 2he Rhe = (21.81)2(11) = 5236 W irwin09_435-490hr.qxd 28-07-2010 12:00 Page 475 SECTION 9.10 A P P L I C AT I O N E X A M P L E S 475 This value, more than 5 kW, is a lot of power! Note, however, that this is the power dissipated under the assumption that the element is connected to the 240-V rms supply 100% of the time. If we manipulate the percentage of the time the element is energized, we can control the average power and thus the average temperature. A fairly standard method for temperature control is shown schematically in Fig. 9.37b, where a standard residential single-phase, three-wire service powers the heating element at 240 V rms and the control circuit at 120 V rms. The temperature switch is connected to three resistors. Each temperature setting produces a different current through the thermostat heater, which is just another resistor. Each switch setting will alter the temperature of the themostat heater that controls the temperature set point. We can calculate the power dissipated in the thermostat heater for each temperature setting. If we let RS be the resistance that corresponds to the switch setting, then PTh = I 2Th RTh 2 120 = a b RTh RTh + RS = 1.3 W for the low setting c = 1.79 W for the medium setting = 4.1 W for the high setting 9.40 We see that the thermostat heater power dissipation is lowest at the low setting and increases as the switch is moved to the high setting. Now let us consider the critical issue in the temperature control system. The thermostat heater is located physically adjacent to the control thermostat (very similar to the ones used to control heat and cooling in your homes where you set the temperature manually). In the dryer, the thermostat heater acts as the desired setting. If the temperature at the thermostat exceeds the setting, then the thermostat switch opens, deenergizing the heating element and allowing it to cool off. When the thermostat determines the temperature is too low, the thermostat switch will close, energizing the element and increasing the temperature. In this way, the thermostat switch opens and closes throughout the drying cycle, maintaining the correct temperature as selected by the temperature switch. Heating element 240 V rms ± – Rhe Figure 9.37 Partial schematics for a residential clothes dryer: (a) the heating element and (b) the control system 11 ⍀ (a) 120 V rms ± – ± – High-limit 'stat switch 1.8 k⍀ Control 'stat switch low med 120 V rms 2.7 k⍀ high 5⍀ 3.5 k⍀ Thermostat heater heat heat Control thermostat (b) heat Rhe High limit thermostat irwin09_435-490hr.qxd 476 28-07-2010 CHAPTER 9 12:00 Page 476 S T E A D Y- S TAT E P O W E R A N A LY S I S Note that all of our calculations have been made for power levels, not temperatures. The exact temperatures of the heating element, the thermostat heater, and the thermostat itself depend on how heat moves about within the dryer—a thermodynamics issue that cannot be addressed with a simple circuit diagram. Finally, note the high limit thermostat and its associated switch. This thermostat is mounted very close to the heating element. If the control thermostat fails, there is no temperature control and we can expect trouble. The high limit thermostat will detect these excessive temperatures and deenergize the heating element. Once the temperature drops, normal operation can resume. Thus, the high limit thermostat is used to protect the dryer and by extension your home. 9.11 The following application-oriented examples illustrate a practical use of the material studied in this chapter. Design Examples DESIGN EXAMPLE 9.27 A light-duty commercial single-phase three-wire 60-Hz circuit serves lighting, heating, and motor loads, as shown in Fig. 9.38a. Lighting and heating loads are essentially pure resistance and, hence, unity power factor (pf), whereas motor loads have lagging pf. We wish to design a balanced configuration for the network and determine its economic viability using the following procedure. a. Compute the phase and neutral currents, as well as the complex power and pf for each source. b. Now move the heating load (panel H) to phase b, as shown in Fig. 9.38b. This is called “balancing” the load. Repeat the analysis of (a). c. Assume that the phase and neutral conductor resistances are each 0.05 ⍀ and have negligible effect on the results of (a). Evaluate the system line losses for (a) and (b). If the loads in question operate 24 hours per day and 365 days per year, at $0.08/kWh, how much energy (and money) is saved by operating in the balanced mode? a. The magnitudes of the rms currents are SOLUTION Figure 9.38 IL = IH = Single-phase three-wire power distribution system. Phase a a Neutral n Im = Ia 10,000 = 41.67 A rms 240 IH Panel L 5 kW In Im IL ± – Panel M 10 kVA pf=0.8 b IH ± – Phase b Ib b (a) Im Panel L n ± – 120 0° V rms Ia a IL Panel L 5 kW ± – 120 0° V rms and P 5000 = = 41.67 A rms V 120 Panel H Ib (b) Panel M irwin09_435-490hr.qxd 28-07-2010 12:00 Page 477 SECTION 9.11 In addition, m = cos-1(0.8) = -36.9° Therefore, Ia = IL + IH + Im = 41.67 / 0° + 41.67 / 0° + 41.67 / -36.9° = 119.4 / -12.1° A rms The currents in the neutral and phase b lines are I n = I L + I H = 83.34 / 0° A rms I b = 41.67 / -36.9° A rms The complex power and power factor for each source are Sa = Van I *a = A120 / 0°BA119.4 / +12.1°B = 14 + j3 kVA pfa = cos (12.1°) = 0.9778 lagging and in a similar manner Sb = Vbn I *b = 4 + j3 kVA pfb = 0.8 lagging b. Under the balanced condition I a = I L + I m = 41.67 / 0° + 41.67 / -36.9° = 79.06 / -18.4° A rms and I b = 79.06 / -18.4° A rms In = 0 Therefore, Sa = Vna I *a = 9 + j3 kVA pfa = 0.9487 lagging and Sb = Vbn I *b = 9 + j3 kVA pfb = 0.9487 lagging c. The power loss in the lines in kW is Ploss = Ra I 2a + Rb I 2b + Rn I 2c = 0.05AI 2a + I 2b + I 2n B兾1000 The total energy loss for a year is Wloss = (24)(365)Ploss = 8760 Ploss and the annual cost is Cost = $0.08 Wloss A comparison of the unbalanced and balanced cases is shown in the following table: Ploss(kW) Wloss(kW-hr) Cost($) UNBALANCED CASE BALANCED CASE 1.147 10,034 804 0.625 5,475 438 Therefore, the annual savings obtained using the balanced configuration is Annual energy savings = 10,048 - 5,475 = 4,573 kWh Annual cost savings = 804 - 438 = $366 DESIGN EXAMPLES 477 irwin09_435-490hr.qxd 478 • CHAPTER 9 12:00 Page 478 S T E A D Y- S TAT E P O W E R A N A LY S I S SUMMARY ■ ■ ■ ■ ■ ■ • 28-07-2010 Instantaneous power If the current and voltage are sinusoidal functions of time, the instantaneous power is equal to a time-independent average value plus a sinusoidal term that has a frequency twice that of the voltage or current. waveform. The rms value of a sinusoidal function is equal to the maximum value of the sinusoid divided by 12 . ■ Power factor Apparent power is defined as the product Vrms Irms . The power factor is defined as the ratio of the average power to the apparent power and is said to be leading when the phase of the current lags the voltage, and lagging when the phase of the current lags the voltage. The power factor of a load with a lagging power factor can be corrected by placing a capacitor in parallel with the load. ■ Complex power The complex power, S, is defined as the product Vrms I *rms . The complex power S can be written as S = P + jQ, where P is the real or average power and Q is the imaginary or quadrature power. Average power P = 1兾2 VI cos Av - i B = 1兾2 VI cos , where is the phase of the impedance. Resistive load P = 1兾2 I 2 R = 1兾2 VI since V and I are in phase. Reactive load P = 1兾2 VI cos A;90°B = 0 Maximum average power transfer To obtain the maximum average power transfer to a load, the load impedance should be chosen equal to the complex conjugate of the Thévenin equivalent impedance representing the remainder of the network. rms or effective value of a periodic waveform The effective, or rms, value of a periodic waveform was introduced as a means of measuring the effectiveness of a source in delivering power to a resistive load. The effective value of a periodic waveform is found by determining the root-mean-square value of the S = I 2Z = I 2R + jI 2X ■ The single-phase three-wire circuit The single-phase three-wire circuit is the one commonly used in households. Large appliances are connected line to line and small appliances and lights are connected line to neutral. ■ Safety Safety must be a primary concern in the design and use of any electrical circuit. The National Electric Code is the primary guide for the construction and operation of electrical systems. PROBLEMS 9.1 The voltage and current at the input of a network are 9.4 Given vs(t) = 100 cos 100t volts, find the average power supplied by the source and the current i2(t) in the network in Fig. P9.4. given by the expressions v(t) = 6 cos t V i2(t) 10 ⍀ i(t) = 4 sin t A 50 mH Determine the average power absorbed by the network. 9.2 The voltage and current at the input of a circuit are given by the expressions vS(t) ± – 1 mF 15 ⍀ v(t) = 170 cos(t + 30°) V i(t) = 5 cos (t + 45°) A Determine the average power absorbed by the circuit. 9.3 Determine the equations for the current and the instantaneous power in the network in Fig. P9.3. Figure P9.4 9.5 Determine the instantaneous power supplid by the source in the circuit in Fig. P9.5. –j2 ⍀ I 4⍀ 4⍀ 12 75° V ± – Figure P9.3 j3 ⍀ 12 60° V + – Figure P9.5 j2 ⍀ 2⍀ irwin09_435-490hr2.qxd 2-08-2010 16:45 Page 479 PROBLEMS 479 9.10 If ig(t) = 0.5 cos 2000t A, find the average power 9.6 Find the instantanous power supplied by the source in the absorbed by each element in the circuit in Fig. P9.10. network in Fig. P9.6. –j1 ⍀ 40 ⍀ 120 ⍀ j1 ⍀ 4⍀ 6 30° A 2⍀ ig(t) 60 mH 12.5 F Figure P9.6 Figure P9.10 9.7 Find the instantaneous travel supplied by the source in the network in Fig. P9.7. –j4 ⍀ 9.11 Find the average power absorbed by the network in Fig. P9.11. j4 ⍀ 1⍀ 4⍀ 4⍀ + 2⍀ 4⍀ 12 0° V 12 30° V 1⍀ + – –j2 ⍀ j2 ⍀ Figure P9.7 –j1 ⍀ Figure P9.11 9.8 Find the instantaneous power supplied by the source in 9.12 Find the average power absorbed by the 2-⍀ resistor in the network in Fig. P9.8. the network in Fig. P9.12 and the total power supplied. 2⍀ –j2 ⍀ j1 ⍀ 4⍀ 2⍀ –j2 ⍀ 6 45° A –j2 ⍀ 4⍀ 6 0° V Figure P9.8 j2 ⍀ Figure P9.12 9.9 Find the average power absorbed by the resistor in the circuit shown in Fig. P9.9 if v1(t) = 10 cos (377t + 60°) V and v2(t) = 20 cos (377t + 120°) V. v2(t) 9.13 Find the average power absorbed by the 2-⍀ resistor in the circuit in Fig. P13 and the total power supplied. v1(t) ± – Figure P9.9 –j1 ⍀ 2⍀ ±– 1⍀ 12 0° A + - Figure P9.13 j2 ⍀ 4⍀ –j2 ⍀ 2⍀ irwin09_435-490hr.qxd 480 28-07-2010 CHAPTER 9 12:00 Page 480 S T E A D Y- S TAT E P O W E R A N A LY S I S 9.14 Find the total average power supplied and the average power –j2 ⍀ 9.19 Given the network in Fig. P9.19, find the power supplied and the average power absorbed by each element. absorbed by each element in the network in Fig. P9.14. 2⍀ ± 4⍀ 2⍀ 4 0° A j2 ⍀ + - 12 0° A 2⍀ VT 4 30° A j3 ⍀ –j1 ⍀ – Figure P9.14 Figure P9.19 9.15 Calculate the average power supplied by the source in the 9.20 Given the network in Fig. P9.20, show that the power supplied by the sources is equal to the power absorbed by the passive elements. circuit in Fig. P9.15. 2⍀ 2H 6⍀ 50 cos5t V –j2 ⍀ + - 1 H 4⍀ 0.05 F j3 ⍀ 0.2 F 5⍀ 3⍀ ± – 4⍀ 2 0° A 6 45° V Figure P9.15 Figure P9.20 9.16 Compute the average power absorbed by each of the elements to the right of the dashed line in the circuit shown in Fig. P9.16. 9.21 Given the network in Fig. P9.21, find the average power supplied to the circuit. j1 ⍀ 1⍀ j1 ⍀ 2⍀ 1⍀ 10 0° V ± – 4 0° V 12 0° A –j1 ⍀ –j2 ⍀ Figure P9.21 Figure P9.16 9.17 Find the average power absorbed by the network shown in 9.22 Calculate the average power absorbed by the 1-⍀ resistor in the network shown in Fig. P9.22. Fig. P9.17. j2 ⍀ j2 ⍀ 2⍀ 4 60° A 6 0° V ± – –j1 ⍀ 2⍀ 1⍀ –j4 ⍀ Figure P9.22 Figure P9.17 9.23 Determine the average power absorbed by the 4-⍀ 9.18 Determine the average power supplied by each source in resistor in the network shown in Fig. P9.23. the network shown in Fig. P9.18. –j2 ⍀ 2⍀ ± 4⍀ j1 ⍀ 4 0° A 2⍀ V ± – 12 0° V – Figure P9.18 Figure P9.23 –j4 ⍀ j2 ⍀ 2 0° A 4⍀ ± – irwin09_435-490hr.qxd 28-07-2010 12:00 Page 481 481 PROBLEMS 9.24 Determine the average power supplied to the network in Fig. P9.24. 9.28 Determine the average power absorbed by the 2- k⍀ output resistor in Fig. P9.28. j1 ⍀ 2 k⍀ 1⍀ 1 k⍀ – ± 2⍀ – j1 ⍀ vS(t)=2 cos t V ± – 1 0° A + vo(t) 2 k⍀ - j1 ⍀ 1⍀ Figure P9.28 Figure P9.24 9.25 Determine the average power absorbed by a 2-⍀ resistor connected at the output terminals of the network shown in Fig. P9.25. Ix 9.29 Determine the impedance Z L for maximum average power transfer and the value of the maximum power transferred to Z L for the circuit shown in Fig. P9.29. j1 ⍀ 2⍀ j1 ⍀ –j2 ⍀ ± – 2Ix ZL Figure P9.29 Figure P9.25 9.30 Determine the impedance Z L for maximum average power 9.26 Find the average power absorbed by the 2-⍀ resistor in the circuit shown in Fig. P9.26. transfer and the value of the maximum average power transferred to Z L for the circuit shown in Fig. P9.30. Ix j2 ⍀ 1⍀ ± – –j3 ⍀ 2⍀ 2Ix 1⍀ 6 0° A ± – 12 0° V 12 0° V Figure P9.26 ± – – j1 ⍀ 12 30° V 4⍀ j1 ⍀ ZL 4 0° A Figure P9.30 9.27 Determine the average power absorbed by the 4- k⍀ resistor in Fig. P9.27. 9.31 Determine the impedance Z L for maximum average power ± – vS(t)=2 cos t V ± – 2 k⍀ transfer and the value of the maximum average power absorbed by the load in the network shown in Fig. P9.31. + vo(t) 1 k⍀ j2 ⍀ 4 k⍀ 4 30° A - Figure P9.27 Figure P9.31 2⍀ 6 0° V ±– – j2 ⍀ ZL irwin09_435-490hr.qxd 482 28-07-2010 CHAPTER 9 12:00 Page 482 S T E A D Y- S TAT E P O W E R A N A LY S I S 9.32 Determine the impedance Z L for maximum average power transfer and the value of the maximum average power absorbed by the load in the network shown in Fig. P9.32. 9.36 Find the value of ZL in Fig. P9.36 for maximum average power transfer to the load. –j4 ⍀ 6 0° V –± 2⍀ + - j1 ⍀ 12 0° V 4⍀ 4⍀ –j1 ⍀ 1⍀ ZL ZL –j4 ⍀ j2 ⍀ 2⍀ Figure P9.32 9.33 Find the value of ZL in Fig. P9.33 for maximum average power transfer to the load. Figure P9.36 9.37 Determine the impedance Z L for maximum average power 6⍀ + transfer and the value of the maximum average power transferred to Z L for the circuit shown in Fig. P9.37. 12 0° V –j1 ⍀ 2⍀ 4⍀ 4 0° V 1⍀ ZL –j4 ⍀ j4 ⍀ 1⍀ ± – j1 ⍀ 4 0° V ZL 2⍀ Figure P9.33 Figure P9.37 9.34 Find the value of ZL in Fig. P9.34 for maximum average power transfer to the load. 9.38 In the network in Fig. P9.38, find Z L for maximum average power transfer and the maximum average power transferred. –j2 ⍀ 2⍀ 12 0° V ZL +j2 ⍀ 4⍀ + - 2⍀ ZL 1⍀ –j2 ⍀ 2 0° A 2⍀ 6 0° V Figure P9.38 Figure P9.34 9.35 Find the value of ZL in Fig. P9.35 for maximum average power transfer to the load. 2⍀ 9.39 In the network in Fig. P9.39, find Z L for maximum average power transfer and the maximum average power transferred. j1 ⍀ –j4 ⍀ 1⍀ j2 ⍀ –j1 ⍀ ZL 12 45° V 12 0° V ± – 4 0° A ZL 4⍀ 1⍀ Figure P9.35 ± – Figure P9.39 j1 ⍀ irwin09_435-490hr.qxd 28-07-2010 12:00 Page 483 483 PROBLEMS 9.40 Determine the impedance Z L for maximum average power transfer and the value of the maximum average power absorbed by the load in the network shown in Fig. P9.40. 9.45 Find the rms value of the waveform shown in Fig. P9.45. v(t) (V) 2⍀ 3 2 ZL 2⍀ 1 ± – –j2 ⍀ 4 0° A 6 0° V 0 1 2 3 4 5 t(s) Figure P9.45 Figure P9.40 9.41 Find the impedance Z L for maximum average power transfer and the value of the maximum average power transferred to Z L for the circuit shown in Fig. P9.41. 12 0° V Fig. P9.46. v(t) (V) ±– j1 ⍀ 9.46 Calculate the rms value of the waveform shown in 3 2 1⍀ 1 1⍀ 2I ZL 0 1 2 3 4 5 6 t(s) 1 – j1 ⍀ –2 I Figure P9.46 Figure P9.41 9.42 Repeat Problem 9.40 for the network in Fig. P9.42. ±– j1 ⍀ 1⍀ Fig. P9.47. 2 Vx –j1 ⍀ v(t) (V) ZL 4 0° A + Vx 9.47 Calculate the rms value of the waveform shown in 3 2 1 1⍀ - 0 2 4 6 8 10 12 14 t(s) Figure P9.47 Figure P9.42 9.43 Compute the rms value of the voltage given by the expression v(t) = 10 + 20 cos (377t + 30°) V. 9.44 Compute the rms value of the voltage given by the waveform shown in Fig. P9.44. 9.48 Calculate the rms value of the waveform in Fig. P9.48. i(t) (A) v(t) (V) 10 2 0 Figure P9.44 1 5 6 10 t(s) 0 Figure P9.48 2 3 4 6 7 t(s) irwin09_435-490hr.qxd 484 28-07-2010 CHAPTER 9 12:00 Page 484 S T E A D Y- S TAT E P O W E R A N A LY S I S 9.50 Calculate the rms value of the waveform shown in Fig. P9.50. 9.49 Calculate the rms value of the waveform in Fig. P9.49. i(t) (A) v(t) (V) 10 6 0 2 3 4 6 7 0 t(s) 2 4 6 8 10 12 14 t(s) Figure P9.50 Figure P9.49 9.51 The current waveform shown in Fig. P9.51 is applied to a 4-⍀ resistor. Calculate the average power dissipated in the resistor. i(t) (A) 1 1 6 t(s) 2 3 4 5 7 8 9 10 –0.5 –1 i(t) 4 k⍀ Figure P9.51 9.52 Find the rms value of the waveform shown in 9.54 Find the rms value of the waveform shown in Fig. P9.52. Fig. P9.54. v(t) (V) v(t) (V) 3 5 2 4 1 3 t(s) 0 1 2 3 2 4 –1 1 –2 0 Figure P9.52 t(s) 1 2 3 4 5 6 7 –1 –2 9.53 Calculate the rms value of the waveform shown in Fig. P9.53. Figure P9.54 9.55 Calculate the rms value of the waveform in Fig. P9.55. v(t) (V) i(t) (A) 2 4 2 0 1 t(s) 1 Figure P9.53 2 3 4 5 6 7 8 –2 Figure P9.55 2 3 4 5 t(s) irwin09_435-490hr2.qxd 2-08-2010 16:45 Page 485 PROBLEMS 9.56 Calculate the rms value of the waveform shown in 485 9.62 The power supply in Fig. P9.59 generates 50 kW the line impedance is 0.095 ⍀. If the load consumes 43 kW and the voltmeter leads 220 V rms, determine the ammeter leading and the power factor of the inductive load. Fig. P9.56. v(t) (V) 4 9.63 A plant consumes 100 kW of power at 0.9 pf lagging. If the load current is 200 A rms, find the load voltage. 0 1 2 3 4 5 6 t(s) 9.64 A plant consumes 20 kW of power from a 240-V rms Figure P9.56 9.57 The current waveform in Fig. P9.57 is flowing through a 5-⍀ resistor. Find the average power absorbed by the resistor. line. If the load power factor is 0.9, what is the angle by which the load voltage leads the load current? What is the load current phasor if the line voltage has a phasor of 240 / 0° V rms? i(t) (A) 9.65 A plant draws 250 A rms from a 240-V rms line to supply 4 a load with 50 kW. What is the power factor of the load? 2 0 2 –2 4 6 8 9.66 The power company must generate 100 kW to supply t(s) an industrial load with 94 kW through a transmission line with 0.09-⍀ resistance. If the load power factor is 0.83 lagging, find the load voltage. –4 Figure P9.57 9.67 A transmission line with impedance of 0.08 + j0.25 ⍀ 9.58 Calculate the rms value of the waveform shown in is used to deliver power to a load. The load is inductive, and the load voltage is 220 / 0° V rms at 60 Hz. If the load requires 12 kW and the real power loss in the line is 560 W, determine the power factor angle of the load. Fig. P9.58. i(t) (A) 1 0 1 2 3 4 5 t(s) 9.68 The power company supplies 80 kW to an industrial load. The load draws 220 A rms from the transmission line. If the load voltage is 440 V rms and the load power factor is 0.8 lagging, find the losses in the transmission line. –2 Figure P9.58 9.69 The power company supplies 40 kW to an industrial load. 9.59 The industrial load in Fig. P9.59 is known to be inductive and consumes 90 kW. The ammeter reading is 260A rms, and the voltmeter reading is 480 V rms. Determine the power factor of the total. The load draws 200 A rms from the transmission line. If the load voltage is 240 V rms and the load power factor is 0.8 lagging, find the losses in the transmission line. 9.70 An industrial load that consumes 40 kW is supplied by Zline ± – Power supply voltage Ammeter Voltmeter Industrial load the power company, through a transmission line with 0.1-⍀ resistance, with 44 kW. If the voltage at the load is 240 V rms, find the power factor at the load. 9.71 A transmission line with impedance 0.1 + j0.2 ⍀ is Figure P9.59 9.60 The industrial load in Fig. P9.59 consumes 110 kW at 0.88 of lagging. The ammeter leads 252A rms. Determine the voltmeter reading. 9.61 The industrial load in Fig. P9.59 consumes 88 kW and the power factor is 0.8 lagging. The power supplied is 96 kW, and the line impedance is 0.1 ⍀. Determine the readings of the ammeter and voltmeter. used to deliver power to a load. The load is capacitive and the load voltage is 240 / 0° V rms at 60 Hz. If the load requires 15 kW and the real power loss in the line is 660 W, determine the input voltage to the line. 9.72 An industrial load operates at 30 kW, 0.8 pf lagging. The load voltage is 240 / 0° V rms. The real and reactive power losses in the transmission-line feeder are 1.8 kW and 2.4 kvar, respectively. Find the impedance of the transmission line and the input voltage to the line. irwin09_435-490hr.qxd 486 28-07-2010 CHAPTER 9 12:00 Page 486 S T E A D Y- S TAT E P O W E R A N A LY S I S 9.73 Find the real and reactive power absorbed by each ± – –j5 ⍀ j3 ⍀ 2⍀ j0.25 ⍀ 0.15 ⍀ ± – 125 0° V rms 95 –10° V rms ± – + 70 kVA 0.92 leading Vs 480 0° V rms - Figure P9.77 Figure P9.73 9.74 Calculate the real and reactive power absorbed by every element (including the sources) in the circuits in Fig. P9.74. 9.78 In the circuit shown in Fig. P9.78, calculate VS, the complex power supplied by the source, and the power factor of the source. j0.2 ⍀ 0.1 ⍀ –j2 ⍀ j4 ⍀ 4⍀ ± – 9.77 In the circuit shown in Fig. P9.77, calculate VS, the complex power supplied by the source, and the power factor of the source. element in the circuit in Fig. P9.73. ± – 125 10° V rms ± – + 25 kW 0.88 leading Vs 150 0° V rms 240 0° V rms - Figure P9.78 9.79 Given the network in Fig. P9.79, find the complex power Figure P9.74 9.75 For the network in Fig. P9.75, the complex power absorbed by the source on the right is 0 + j1582.5 VA. Find the value of R and the unknown element and its value if f = 60 Hz . (If the element is a capacitor, give its capacitance; if the element is an inductor, give its inductance.) ± – ± – 120 –10° V rms j0.2 ⍀ 0.05 ⍀ ± 20 kW 0.7 leading ± V S – 12 kVA 0.9 lagging 240 0° V rms – Figure P9.79 Unknown element R supplied by the source, the power factor of the source, and the voltage vs(t). The frequency is 60 Hz. 150 0° V rms 9.80 Find the complex power supplied by the source, the power factor of the source, and VS, (t) if f = 60 Hz in Fig. P9.80 0.15 ⍀ j0.3 ⍀ ± Figure P9.75 40 kW 0.84 lagging ± V S – 40 kVA 0.9 leading – 9.76 The source in the circuit in Fig. P9.76 supplies 40 kW at a power factor of 0.9 lagging. The real and reactive losses of the transmission-line feeder are 1.6 kW and 2.1kvar, respectively. Calculate R, X, and the load voltage. R Figure P9.80 9.81 Use Kirchhoff’s laws to compute the source voltage of the network shown in Fig. P9.81. jX 0.09 ⍀ + ± – 480 0° V rms Figure P9.76 480 0° V rms VL - Load VS ± – j0.25 ⍀ ± 24 kW 0.85 pf lagging 220 0° V rms – Figure P9.81 36 kW 0.78 pf lagging irwin09_435-490hr.qxd 28-07-2010 12:00 Page 487 PROBLEMS 9.83 Given the network in Fig. P9.83, determine the input 9.82 Given the network in Fig. P9.82, determine the input voltage VS. voltage VS. 0.1 ⍀ 487 j0.3 ⍀ 36 kW 0.82 pf lagging VS ± – 48 kW 0.88 pf lagging j0.1 ⍀ 0.1 ⍀ + 240 0° V rms + 30 kVA 0.9 pf lagging VS ± – 40 kW 0.795 pf lagging 240 0° V rms - - Figure P9.83 Figure P9.82 9.84 Given the circuit in Fig. P9.84, find the complex power supplied by the source and the source power factor. If f = 60 Hz, find vs(t). 0.1 ⍀ j0.2 ⍀ 30 kW 0.8 leading ± V S – 20 kVA 0.9 lagging 10 kW 0.8 lagging ± 480 0° V rms – Figure P9.84 9.85 Given the network in Fig. P9.85, compute the input source voltage and the input power factor. j0.2 ⍀ 0.08 ⍀ 0.01 ⍀ 60 kW 0.86 pf lagging VS ± – j0.05 ⍀ + 20 kW 0.8 pf lagging 220 0° V rms - Figure P9.85 9.86 Given the network in Fig. P9.86, compute the input source voltage and the input power factor. 0.08 ⍀ j0.2 ⍀ 0.01 ⍀ 9.89 A particular load has a pf of 0.8 lagging. The power delivered to the load is 40 kW from a 270-V rms 60-Hz line. What value of capacitance placed in parallel with the load will raise the pf to 0.9 lagging? j0.05 ⍀ + ± VS – 40 kW 0.86 pf lagging 18 kW 0.8 pf lagging 9.90 A particular load has a pf of 0.8 lagging. The power 220 0° V rms - Figure P9.86 9.87 What value of capacitance must be placed in parallel with the 18-kW load in Problem 9.86 to raise the power factor of this load to 0.9 lagging? 9.88 An industrial load consumes 44 kW at 0.82 pf lagging from a 240 / 0° - V rms 60-Hz line. A bank of capacitors totaling 600 F is available. If these capacitors are placed in parallel with the load, what is the new power factor of the total load? delivered to the load is 40 kW from a 220-V rms 60-Hz line. What value of capacitance placed in parallel with the load will raise the pf to 0.9 lagging? 9.91 An industrial load is supplied through a transmission line that has a line impedance of 0.1 + j0.2 ⍀. The 60-Hz line voltage at the load is 480 / 0° V rms. The load consumes 124 kW at 0.75 pf lagging. What value of capacitance when placed in parallel with the load will change the power factor to 0.9 lagging? 9.92 The 60-Hz line voltage for a 60-kW, 0.76-pf lagging industrial load is 240 / 0° V rms. Find the value of capacitance that when placed in parallel with the load will raise the power factor to 0.9 lagging. irwin09_435-490hr3.qxd 488 9-08-2010 CHAPTER 9 16:19 Page 488 S T E A D Y- S TAT E P O W E R A N A LY S I S 9.93 A plant consumes 60 kW at a power factor of 0.75 lag- 9.98 Determine the value of capacitance in Fig. P9.98 that ging from a 240-V rms 60-Hz line. Determine the value of the capacitor that when placed in parallel with the load will change the load power factor to 0.9 lagging. must be connected in parallel with the load so that the power factor of the combined load and capacitor is unity. Calculate the complex power supplied by the source after the power factor has been corrected to unity. The frequency f = 60 Hz 9.94 A bank of induction motors consumes 36 kW at 0.78 pf lagging from a 60-Hz 240 / 0° - V rms line. If 200 F of capacitors are placed in parallel with the load, what is the new power factor of the total load? 0.15 ⍀ ± – j0.3 ⍀ 40 kVA 0.84 lagging Vs + 480 0° V rms - 9.95 Calculate the value of capacitance in Fig. P9.95 that must be connected in parallel with the load to correct the source power factor with the load to correct the source power factor to 0.94 lagging. The frequncy f = 60 Hz. 9.99 A 5-kW load operates at 60 Hz, 240-V rms and has a 30 kW 0.75 lagging ± – 240 0° V rms Figure P9.98 power factor of 0.866 lagging. We wish to create a power factor of at least 0.975 lagging using a single capacitor. Can this requirement be met using a single capacitor from Table 6.1? Figure P9.95 9.96 In the circuit in Fig. P9.96, a load is modeled by an impedance of 4 + j4 E. Determine the value of capacitance that must be connected in parallel with the load to correct the power factor of the combined load and capacitor to 0.95 lagging. The frequncy f = 60 Hz. 9.100 A 5.1-kW household range is designed to operate on a 240-V rms sinusoidal voltage, as shown in Fig. P9.100a. However, the electrician has mistakenly connected the range to 120 V rms, as shown in Fig. P9.100b. What is the effect of this error? j0.8 ⍀ 4⍀ a + Vs 4⍀ ± – 120 0° V rms j4 ⍀ A ± – 120 0° V rms - Range 5100 W Figure P9.96 ± – 120 0° V rms 9.97 In the circuit in Fig. P9.97, a load is modeled by an impedance of 4 + j4 ⍀. Determine the value of capaci- b tance which must be connected in parallel in with the load to correct the power factor of the combined load and capacitor to 0.95 lagging. The frequency f = 60 Hz. 4⍀ 120 0° V rms ± – (a) a j0.8 ⍀ 4⍀ j4 ⍀ B A ± – 120 0° V rms Range n N (b) Figure P9.97 Figure P9.100 irwin09_435-490hr.qxd 28-07-2010 12:00 Page 489 PROBLEMS 9.101 To test a light socket, a woman, while standing on cushions that insulate her from the ground, sticks her finger into the socket, as shown in Fig. P9.101. The tip of her finger makes contact with one side of the line, and the side of her finger makes contact with the other side of the line. Assuming that any portion of a limb has a resistance of 95 ⍀, is there any current in the body? Is there any current in the vicinity of the heart? 489 9.103 A single-phase three-wire 60-Hz circuit serves three loads, as shown in Fig. P9.103. Determine I aA, I nN, Ic, and the energy use over a 24-hour period in kilowatthours. IaA a ± – 120 0° V rms ± – 120 0° V rms Ic 100 W InN n A N 1 kVA pf=0.9 lagging 100 W b Figure P9.103 9.104 A number of 120-V rms household fixtures are to be Figure P9.101 9.102 An inexperienced mechanic is installing a 12-V battery in a car. The negative terminal has been connected. He is currently tightening the bolts on the positive terminal. With a tight grip on the wrench, he turns it so that the gold ring on his finger makes contact with the frame of the car. This situation is modeled in Fig. P9.102, where we assume that the resistance of the wrench is negligible and the resistance of the contact is as follows: R1 = Rbolt to wrench = 0.012 ⍀ R2 = Rwrench to ring = 0.012 ⍀ R3 = Rring = 0.012 ⍀ R4 = Rring to frame = 0.012 ⍀ What power is quickly dissipated in the gold ring, and what is the impact of this power dissipation? R1 ± – R2 R3 12 V R4 Figure P9.102 used to provide lighting for a large room. The total lighting load is 8 kW. The National Electric Code requires that no circuit breaker be larger than 20 A rms with a 25% safety margin. Determine the number of identical branch circuits needed for this requirement. 9.105 A man and his son are flying a kite. The kite becomes entangled in a 7200-V rms power line close to a power pole. The man crawls up the pole to remove the kite. While trying to remove the kite, the man accidentally touches the 7200-V rms line. Assuming that the power pole is well grounded, what is the potential current through the man’s body? irwin09_435-490hr.qxd 490 28-07-2010 CHAPTER 9 12:00 Page 490 S T E A D Y- S TAT E P O W E R A N A LY S I S TYPICAL PROBLEMS FOUND ON THE FE EXAM • 9PFE-1 An industrial load consumes 120 kW at 0.707 pf lagging and is connected to a 480 / 0°V rms 60-Hz line. 9PFE-4 An rms-reading voltmeter is connected to the output of the op-amp shown in Fig. 9PFE-4. Determine the meter reading. Determine the value of the capacitor that, when connected in parallel with the load, will raise the power factor to 0.95 lagging. a. 3 V b. 5.2 V a. 642 F c. 4.24 V b. 763 F d. 2 V c. 928 F d. 471 F 36 k⍀ 9PFE-2 Determine the rms value of the following waveform. 12 k⍀ a. 2.33 V – ± b. 1 V c. 3.25 V ± – 1.414 cos t V d. 1.22 V rms-reading voltmeter v(t) (V) 2 Figure 9PFE-4 1 0 1 2 3 4 5 9PFE-5 Determine the average power delivered to the resistor t(s) in Fig. 9PFE-5a if the current waveform is shown in Fig. 9PFE-5b. Figure 9PFE-2 a. 18.78 W 9PFE-3 Find the impedance Z L in the network in Fig. 9PFE-3 b. 8.64 W for maximum power transfer. c. 2.82 W a. 0.8 + j2.4 ⍀ d. 10.91 W b. 0.4 - j1.2 ⍀ c. 0.2 + j1.4 ⍀ i(t) (A) d. 0.3 - j1.6 ⍀ 2 i(t) 12 0° V ±– 2⍀ –j1 ⍀ Figure 9PFE-3 ZL 1 0 1 2 –2 (a) 2 0° A 4⍀ j2 ⍀ Figure 9PFE-5 (b) 3 4 t(s) irwin10_491-540hr.qxd 28-07-2010 12:10 Page 491 10 C H A PT E R MAGNETICALLY COUPLED NETWORKS THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Understand the concepts of mutual inductance, coefficient of coupling, and turns ratio ■ Know how to calculate voltages and currents in circuits containing mutual inductance ■ Know how to calculate voltages and currents in circuits containing ideal transformers Fudan Li/iStockphoto M Magnetic Levitation Train The Pony Express is to today’s wheel noises; no exhaust fumes; and amazing speed. The only electronic mail as the stagecoach of the Old West is to which drawbacks are higher costs and lack of compatibility with exist- modern miracle of land transportation? Did you guess a ing infrastructure. speeding locomotive? Very good, but a better answer is the This chapter introduces a central concept of MagLev tech- magnetic levitation train—abbreviated MagLev. These passen- nology: the magnetic effects of mutual coupling in circuits ger trains—the world’s fastest—float on a magnetic field resulting from changing currents. Voltages are induced in near- about 10 mm above the guideway, propelled by a linear by branches due to these magnetic fields—this coupling is induction motor. Shanghai Pudong International Airport has a quantified by mutual inductance terms. Transformers that step MagLev that reaches speeds over 300 mph, 60% faster than up or step down voltages and currents from their primary to the famed Bullet Trains of Japan. It gives the smoothest ride secondary sides are useful in several ways, such as transmit- ever, coasting along under magnetic power. ting power over long distances or for impedance matching in MagLev trains are used primarily when two large cities are acoustical equipment. Requiring extremely strong magnetic to be connected for passenger service. MagLev systems have fields, MagLev trains depend on these same mutual coupling several advantages: lower maintenance because there are no effects for high-speed transportation. It’s not magic; it’s elec- moving parts; no friction, only air resistance; an absence of tromagnetics. 491 irwin10_491-540hr.qxd 492 28-07-2010 CHAPTER 10 10.1 12:10 Page 492 M A G N E T I C A L LY C O U P L E D N E T W O R K S As we introduce this subject, we feel compelled to remind the reader, once again, that in our analyses we assume that we are dealing with “ideal” elements. For example, we ignore the resistance of the coil used to make an inductor and any stray capacitance that might exist. This approach is especially important in our discussion of mutual inductance because an exact analysis of this topic is quite involved. As is our practice, we will treat the subject in a straightforward manner and ignore issues beyond the scope of this book that only serve to complicate the presentation. To begin our discussion of mutual inductance, we will recall two important laws: Ampère’s law and Faraday’s law. Ampère’s law predicts that the flow of electric current will create a magnetic field. If the field links an electric circuit, and that field is time-varying, Faraday’s law predicts the creation of a voltage within the linked circuit. Although this occurs to some extent in all circuits, the effect is magnified in coils because the circuit geometry amplifies the linkage effect. With these ideas in mind, consider the ideal situation in Fig. 10.1 in which a current i flows in an N-turn coil and produces a magnetic field, represented by magnetic flux . The flux linkage for this coil is Mutual Inductance = N Figure 10.1 Magnetic flux linking an N-turn coil. 10.1 + i v N turns - For the linear systems that we are studying in this textbook, the flux linkage and current are related by = Li 10.2 The constant of proportionality between the flux linkage and current is the inductance, which we studied in Chapter 6. Eqs. (10.1) and (10.2) can be utilized to express the magnetic flux in terms of the current: = L i N 10.3 According to Faraday’s law, the voltage induced in the coil is related to the time rate of change of the flux linkage : v = i v = - Figure 10.2 An ideal inductor. 10.4 Let’s substitute Eq. (10.2) into Eq. (10.4) and use the chain rule to take the derivative: + v d dt L d di dL d = (Li) = L + i dt dt dt dt 10.5 We will not allow our inductances to vary with time, so Eq. (10.5) reduces to the defining equation for the ideal inductor, as shown in Fig. 10.2: v = L di dt 10.6 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 493 SECTION 10.1 M U T U A L I N D U C TA N C E Figure 10.3 + i1 v1 Two coils magnetically coupled. + N1 493 N2 v2 - - Note that the voltage and current in this figure satisfy the passive sign convention. Equation (10.6) tells us that a current i flowing through a coil produces a voltage v across that coil. Now let’s suppose that a second coil with N2 turns is moved close enough to an N1-turn coil such that the magnetic flux produced by current i1 links the second coil. No current flows in the second coil as shown in Fig. 10.3. By Faraday’s law, a voltage v2 will be induced because the magnetic flux links the second coil. The flux linkage for coil 1 is 1 = N1 = L1 i1 10.7 d1 di 1 = L1 . We have been referring to L1 dt dt as the inductance. In multiple coil systems, we will refer to L1 as the self-inductance of coil 1. The flux linkage for coil 2 is 2 = N2 , and from Faraday’s law, the voltage v2 is given as Current flowing in coil 1 produces a voltage v1 = v2 = L1 d2 N2 di 1 di 1 d d = AN2 B = a N2 a i1 b b = L1 = L21 dt dt dt N1 N1 dt dt 10.8 Note that the voltage v2 is directly proportional to the time rate of change of i1 . The constant of proportionality, L21 , is defined as the mutual inductance and is given in units of henrys. We will say that the coils in Fig. 10.3 are magnetically coupled. Let’s connect a current source to the terminals of coil 2 as shown in Fig. 10.4. Both currents contribute to the magnetic flux . For the coil configuration and current directions shown in this figure, the flux linkages for each coil are 1 = L1 i1 + L12 i2 10.9 2 = L21 i1 + L2 i2 10.10 Applying Faraday’s law, v1 = d1 di 1 di 2 = L1 + L12 dt dt dt 10.11 v2 = d2 di 1 di 2 = L21 + L2 dt dt dt 10.12 Figure 10.4 + i1 v1 - + N1 N2 v2 - i2 Two magnetically coupled coils driven by current sources. irwin10_491-540hr.qxd 494 28-07-2010 CHAPTER 10 12:10 Page 494 M A G N E T I C A L LY C O U P L E D N E T W O R K S Since we have limited our study to linear systems, L12 = L21 = M, where M is the symbol for mutual inductance. From Eqs. (10.11) and (10.12), we can see that the voltage across each coil is composed of two terms: a “self term” due to current flowing in that coil and a “mutual term” due to current flowing in the other coil. If the direction of i2 in Fig. 10.4 is reversed, Eqs. (10.9) through (10.12) become 1 = L1 i1 - Mi2 10.13 2 = -Mi1 + L2 i2 10.14 v1 = d1 di 1 di 2 = L1 - M dt dt dt 10.15 v2 = d2 di 2 di 1 = -M + L2 dt dt dt 10.16 Eqs. (10.13)–(10.16) can also be obtained from the circuit in Fig. 10.5. Note that coil 2 in this figure has a different winding arrangement as compared to coil 2 in Fig. 10.4. Our circuit diagrams will become quite complex if we have to include details of the winding configuration. The use of the dot convention permits us to maintain these details while simplifying our circuit diagrams. Fig. 10.6a is the circuit diagram for the magnetically coupled coils of Fig. 10.4. The coils are represented by two coupled inductors with selfinductances L1 and L2 and mutual inductance M. Recall that the voltage across each coil consists of two terms: a self term due to current flowing in that coil and a mutual term due to current flowing in the other coil. The self term is the same voltage that we discussed in an earlier chapter. The mutual term results from current flowing in the other coupled coil. Figure 10.5 Magnetically coupled coils with different winding configuration. + i1 v1 + N1 N2 - i2 v2 - Figure 10.6 M Circuit diagrams for magnetically coupled coils. + i1 v1 + + di1 di2 L1 ––– +M ––– L1 dt dt - + di2 di1 L2 M ––– + L2 ––– v2 dt dt - - i2 - (a) M + i1 v1 - + + di1 di2 L1 ––– -M ––– L1 dt dt + di2 di1 L2 –M ––– + L2 ––– v2 dt dt - (b) - i2 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 495 SECTION 10.1 M U T U A L I N D U C TA N C E Figure 10.7 M + i1 di1 L1 ––– dt - + + + di2 L2 L2 ––– dt di2 L1 M ––– dt 495 - di1 M ––– dt - - + - di2 L2 ––– dt di1 M ––– dt - + i2 Circuit diagrams for magnetically coupled coils showing self and mutual voltage terms. (a) M - i1 + di1 di2 L1 M ––– L1 ––– dt dt + L2 - i2 (b) In Fig. 10.6a, the mutual terms are positive when both currents enter the dots. The opposite is true when one current enters a dot and the other current leaves a dot, as seen in Fig. 10.6b. Let’s use this observation to develop a general procedure for writing circuit equations for magnetically coupled inductors. Fig. 10.7a is the same diagram as Fig. 10.6a except that the voltage across the inductors is broken into the self term and the mutual term. The polarity of the self terms—L1 di 1兾dt and L2 di 2兾dt—are given by the passive sign convention used extensively throughout this text. These terms would be present even if the coils were not magnetically coupled. The mutual terms in Fig. 10.7a have the same polarity as the self terms. Note that both currents are entering the dots in Fig. 10.7a. The opposite is true in Fig. 10.7b. The self terms have the same polarity as before; however, the polarities for the mutual terms are different from those in Fig. 10.7a. We can now make a general statement: When a current is defined to enter the dotted terminal of a coil, it produces a voltage in the coupled coil which is positive at the dotted terminal. Similarly, when a current is defined to enter the undotted terminal of a coil, it produces a voltage in the coupled coil which is positive at the undotted terminal. Let’s illustrate the use of this statement through some examples. Problem-Solving Strategy Step 1. Assign mesh currents. It is usually much easier to write mesh equations for a circuit containing magnetically coupled inductors than nodal equations. Step 2. Write mesh equations by appling KVL. If a defined current enters the dotted terminal on one coil, it produces a voltage in the other coil that is positive at the dotted terminal. If a defined current enters the undotted terminal on one coil, it produces a voltage in the other coil that is positive at the undotted terminal. Step 3. Solve for the mesh currents. Magnetically Coupled Inductors irwin10_491-540hr.qxd 496 28-07-2010 CHAPTER 10 EXAMPLE 10.1 SOLUTION 12:10 Page 496 M A G N E T I C A L LY C O U P L E D N E T W O R K S Determine the equations for v1(t) and v2(t) in the circuit shown in Fig. 10.8a. The different voltage terms for the circuit are shown on the circuit diagram in Fig. 10.8b. The polarity of the self terms is given by the passive sign convention. For both coils, the defined currents are entering the undotted terminals on both coils. As a result, the polarity of the voltages produced by these currents is positive at the undotted terminal of the other coils. The equations for v1(t) and v2(t) are v1(t) = -L1 v2(t) = L2 Figure 10.8a di 2 di 1 + M dt dt M i1(t) Circuit used in Example 10.1. di 1 di 2 - M dt dt i2(t) - + L1 v1(t) L2 v2(t) + - Figure 10.8b Circuit showing self and mutual voltage terms. M i1(t) - + + + di2 di1 L1 v1(t) M ––– L1 ––– dt dt + EXAMPLE 10.2 SOLUTION - + + di2 di1 L2 L2 ––– v2(t) M ––– dt dt - - - - Write mesh equations for the circuit of Fig. 10.9a using the assigned mesh currents. The circuit in Fig. 10.9b shows the voltage terms for mesh 1. The polarity of the self terms for L1 and L2 is determined by the passive sign convention. The current Ai2 - i1 B enters the dotted terminal of inductor L2 . This current produces the mutual term shown across inductor L1 . Current i1 enters the dotted terminal of L1 and produces a voltage across L2 that is positive at its dotted terminal. The equation for this mesh is v1(t) = R1 i1(t) + L1 Figure 10.9a Circuit used in Example 10.2. i2(t) di 1 di 1 d d + M Ai2 - i1 B + L2 Ai1 - i2 B - M dt dt dt dt L1 R1 v1(t) ± – i1(t) M L2 i2(t) R2 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 497 SECTION 10.1 497 M U T U A L I N D U C TA N C E The voltage terms for the second mesh are shown in Fig. 10.9c. The equation for mesh 2 is R2 i2(t) + L2 di 1 d Ai - i1 B + M = 0 dt 2 dt d dt di1 + L ––– 1 dt + M ––– (i2-i1) - v1(t) ± – - i2(t) + di1 d M ––– (i -i ) dt L2 L2 ––– dt 1 2 M i1(t) + - - - R2 i2(t) di1 d M ––– (i -i ) dt L2 L2 ––– dt 2 1 M i1(t) Circuit showing voltage terms for mesh 1. Figure 10.9c L1 R1 v1(t) ± – Figure 10.9b L1 R1 + Circuit showing voltage terms for mesh 2. R2 + Learning Assessment E10.1 Write the equations for v1(t) and v2(t) in the circuit in Fig. E10.1. ANSWER: M i1(t) v1(t) = L1 i2(t) - di 1(t) + L1 v1(t) L2 + v2(t) = -L2 v2(t) dt +M di 2(t) dt di 2(t) -M dt di 1(t) - Figure E10.1 Assume that the coupled circuit in Fig. 10.10 is excited with a sinusoidal source. The voltages will be of the form V1 ejt and V2 ejt, and the currents will be of the form I 1 ejt and I 2 ejt, where V1 , V2 , I1 , and I2 are phasors. Substituting these voltages and currents into Eqs. (10.11) and (10.12), and using the fact that L12 = L21 = M, we obtain V1 = jL1 I 1 + jMI 2 10.17 V2 = jL2 I 2 + jMI 1 i1(t) i2(t) M - v1(t) + L1 Figure 10.10 + L2 v2(t) - ; Mutually coupled coils. dt . irwin10_491-540hr.qxd 498 28-07-2010 CHAPTER 10 12:10 Page 498 M A G N E T I C A L LY C O U P L E D N E T W O R K S The model of the coupled circuit in the frequency domain is identical to that in the time domain except for the way the elements and variables are labeled. The sign on the mutual terms is handled in the same manner as is done in the time domain. The two mutually coupled coils in Fig. 10.11a can be interconnected in four possible ways. We wish to determine the equivalent inductance of each of the four possible interconnections. EXAMPLE 10.3 SOLUTION Case 1 is shown in Fig. 10.11b. In this case V = jL1 I + jMI + jL2 I + jMI = jLeq I where Leq = L1 + L2 + 2M. Case 2 is shown in Fig. 10.11c. Using KVL, we obtain V = jL1 I - jMI + jL2 I - jMI = jLeq I where Leq = L1 + L2 - 2M. Case 3 is shown in Fig. 10.11d and redrawn in Fig. 10.11e. The two KVL equations are V = jL1 I 1 + jMI 2 V = jM I 1 + jL2I 2 Solving these equations for I1 and I2 yields I1 = I2 = Using KCL gives us VAL2 - MB jAL1 L2 - M 2 B VAL1 - MB jAL1 L2 - M 2 B I = I1 + I2 = VAL1 + L2 - 2MB where Leq = jAL1 L2 - M B 2 = V jLeq L1 L2 - M 2 L1 + L2 - 2M Case 4 is shown in Fig. 10.11f. The voltage equations in this case will be the same as those in case 3 except that the signs of the mutual terms will be negative. Therefore, Figure 10.11 Leq = Circuits used in Example 10.3. jM jL1 1 2 L1 L2 - M 2 L1 + L2 + 2M 3 4 1 2 I (a) jM jL1 jL2 jL2 3 V ±– (b) 4 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 499 SECTION 10.1 jM jL1 1 2 I jL2 3 jM jL1 4 1 V ±– 2 I1 I2 I V ±– (c) I M U T U A L I N D U C TA N C E 499 jL2 3 4 (d) 1 I 3 1 4 jM V ± – jL1 V ± – jL2 I1 2 jM jL1 jL2 I2 2 4 (e) 3 (f) Figure 10.11 (continued) EXAMPLE We wish to determine the output voltage Vo in the circuit in Fig. 10.12. SOLUTION The two KVL equations for the network are 10.4 (2 + j4)I 1 - j2I 2 = 24 / 30° -j2I 1 + (2 + j6 - j2)I 2 = 0 Solving the equations yields I 2 = 2.68 / 3.43° A Therefore, Vo = 2I 2 = 5.36 / 3.43° V 2⍀ j2 ⍀ –j2 ⍀ Figure 10.12 + 24 30° V ± – I1 j4 ⍀ j6 ⍀ I2 2⍀ Vo - Example of a magnetically coupled circuit. irwin10_491-540hr.qxd 500 28-07-2010 CHAPTER 10 12:10 Page 500 M A G N E T I C A L LY C O U P L E D N E T W O R K S Let us now consider a more complicated example involving mutual inductance. EXAMPLE Consider the circuit in Fig. 10.13. We wish to write the mesh equations for this network. 10.5 SOLUTION Because of the multiple currents that are present in the coupled inductors, we must be very careful in writing the circuit equations. The mesh equations for the phasor network are I 1 R1 + jL1 AI 1 - I 2 B + jMAI 2 - I 3 B + 1 AI - I 2 B = V jC1 1 1 AI - I 1 B + jL1 AI 2 - I 1 B + jMAI 3 - I 2 B + R2 I 2 jC1 2 + jL2 AI 2 - I 3 B + jMAI 1 - I 2 B + R3 AI 2 - I 3 B = 0 R3 AI 3 - I 2 B + jL2 AI 3 - I 2 B + jMAI 2 - I 1 B + 1 I + R4 I 3 = 0 jC2 3 which can be rewritten in the form aR1 + jL1 + 1 1 b I 1 - ajL1 + - jMb I 2 jC1 jC1 - jMI 3 = V - a jL1 + +a 1 - jM b I 1 jC1 1 + jL1 + R2 + jL2 + R3 - j2M b I 2 jC1 - AjL2 + R3 - jMBI 3 = 0 -jMI 1 - AR3 + jL2 - jMBI 2 + a R3 + jL2 + 1 + R4 b I 3 = 0 jC2 Note the symmetrical form of these equations. 1 ––––– jC2 Figure 10.13 R1 Example of a magnetically coupled circuit. R2 jM jL1 V ± – I1 jL2 I2 1 ––––– jC1 I3 R3 R4 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 501 SECTION 10.1 501 M U T U A L I N D U C TA N C E Learning Assessments E10.2 Find the currents I1 and I2 and the output voltage Vo in the network in Fig. E10.2. 24 0° V j1 ⍀ ANSWER: I 1 = +4.29 / 137.2° A; 2⍀ ±– I 2 = 0.96 / -16.26° A; + Vo = 3.84 / -106.26° V. 4⍀ j4 ⍀ I1 j8 ⍀ I2 –j4 ⍀ Vo - Figure E10.2 ANSWER: AR1 + jL1 + R2 BI 1 E10.3 Write the KVL equations in standard form for the network in Fig. E10.3. jM jL1 - AR2 + jMBI 2 = -V1 ; - AR2 + jMBI 1 jL2 + AR2 + jL2 + R3 BI 2 R2 I1 R1 I2 = V1 . R3 V1 ± – Figure E10.3 E10.4 Find Vo in Fig. E10.4. j 1.5 ⍀ 2⍀ –j2 ⍀ ANSWER: Vo = 11.2 / 53.5° V. j4 ⍀ + 2⍀ 24 0° V + – j2 ⍀ Vo j2 ⍀ 10 30° V + – – Figure E10.4 4⍀ –j 3 ⍀ ANSWER: Vo = 32.8 / –12.14° V. E10.5 Find Vo in Fig. E10.5. –j3 ⍀ 4⍀ j4 ⍀ + j2 ⍀ 36 0° V + – 4⍀ Vo j2 ⍀ 24 10° V + – –j4 ⍀ – Figure E10.5 j4 ⍀ irwin10_491-540hr.qxd 502 28-07-2010 CHAPTER 10 12:10 Page 502 M A G N E T I C A L LY C O U P L E D N E T W O R K S ANSWER: Vo = 11.4 / 0.334° V. E10.6 Find Vo in Fig. E10.6. 2⍀ j1 ⍀ 4⍀ 4⍀ + j3 ⍀ –j 1 ⍀ j3 ⍀ 3 30° V 20 30° V + – Vo 2⍀ 4⍀ + – 10 0° V j2 ⍀ – Figure E10.6 EXAMPLE 10.6 Given the network in Fig. 10.14 with the parameters Z S = 3 + j1 ⍀, jL1 = j2 ⍀, jL2 = j2 ⍀, jM = j1 ⍀, and Z L = 1 - j1 ⍀, determine the impedance seen by the source VS . SOLUTION The mesh equations for the network are VS = AZ S + jL1 BI 1 - jMI 2 0 = -jMI 1 + AjL2 + Z L BI 2 If we now define Z 1 1 = Z S + jL1 and Z 2 2 = jL2 + Z L , then the second equation yields I2 = jM I Z22 1 If this secondary mesh equation is substituted into the primary mesh equation, we obtain VS = Z 1 1 I 1 + 2M 2 I Z22 1 jM ZS VS ± – Figure 10.14 Circuit employed in Example 10.6. I1 I2 jL1 jL2 ZL and therefore VS 2M 2 = Z1 1 + I1 Z22 which is the impedance seen by VS . Note that the mutual term is squared, and therefore the impedance is independent of the location of the dots. irwin10_491-540hr.qxd 28-07-2010 12:10 Page 503 SECTION 10.2 E N E R G Y A N A LY S I S 503 Using the values of the circuit parameters, we find that VS 1 = (3 + j1 + j2) + I1 j2 + 1 - j1 = 3 + j3 + 0.5 - j0.5 = 3.5 + j2.5 ⍀ Learning Assessment E10.7 Find the impedance seen by the source in the circuit in Fig. E10.7. ANSWER: Z S = 2.25 / 20.9° ⍀. j1 ⍀ – j1 ⍀ 2⍀ 120 0° V 2⍀ ± – j2 ⍀ j2 ⍀ –j2 ⍀ j1⍀ Figure E10.7 We now perform an energy analysis on a pair of mutually coupled inductors, which will yield some interesting relationships for the circuit elements. Our analysis will involve the performance of an experiment on the network shown in Fig. 10.15. Before beginning the experiment, we set all voltages and currents in the circuit equal to zero. Once the circuit is quiescent, we begin by letting the current i1(t) increase from zero to some value I1 with the right-side terminals open circuited. Since the right-side terminals are open, i2(t) = 0, and therefore the power entering these terminals is zero. The instantaneous power entering the left-side terminals is p(t) = v1(t)i1(t) = c L1 di 1(t) dt 10.2 Energy Analysis d i1(t) The energy stored within the coupled circuit at t1 when i1(t) = I1 is then t1 I1 v1(t)i1(t) dt = 30 L1i1(t) di 1(t) = 30 1 L I 21 2 1 Continuing our experiment, starting at time t1 , we let the current i2(t) increase from zero to some value I2 at time t2 while holding i1(t) constant at I1 . The energy delivered through the right-side terminals is t2 3t1 I2 v2(t)i2(t) dt = 30 L2 i2(t) di 2(t) = 1 L I 22 2 2 Figure 10.15 M i1(t) i2(t) + v1(t) - Magnetically coupled circuit. + L1 L2 v2(t) - irwin10_491-540hr.qxd 504 28-07-2010 CHAPTER 10 12:10 Page 504 M A G N E T I C A L LY C O U P L E D N E T W O R K S However, during the interval t1 to t2 the voltage v1(t) is di 1(t) v1(t) = L1 dt + M di 2(t) dt Since i1(t) is a constant I1 , the energy delivered through the left-side terminals is t2 3t1 t2 v1(t)i1(t) dt = 3t1 M di 2(t) dt I2 I1 dt = MI1 30 di2(t) = MI1 I2 Therefore, the total energy stored in the network for t 7 t2 is w = 1 1 L1 I 21 + L2 I 22 + MI1 I2 2 2 10.18 We could, of course, repeat our entire experiment with either the dot on L1 or L2 , but not both, reversed, and in this case the sign on the mutual inductance term would be negative, producing w = 1 1 L1 I 21 + L2 I 22 - MI1 I2 2 2 It is very important for the reader to realize that in our derivation of the preceding equation, by means of the experiment, the values I1 and I2 could have been any values at any time; therefore, the energy stored in the magnetically coupled inductors at any instant of time is given by the expression 1 1 2 2 L C i (t)D + L2 C i2(t)D ; Mi1(t)i2(t) 2 1 1 2 w(t) = 10.19 The two coupled inductors represent a passive network, and therefore, the energy stored within this network must be nonnegative for any values of the inductances and currents. The equation for the instantaneous energy stored in the magnetic circuit can be written as w(t) = 1 1 L i21 + L2 i22 ; Mi1 i2 2 1 2 Adding and subtracting the term 1兾2 AM 2兾L2 Bi21 and rearranging the equation yields w(t) = 2 1 M2 2 1 M a L1 b i1 + L2 a i2 + i1 b 2 L2 2 L2 From this expression we recognize that the instantaneous energy stored will be nonnegative if M ⱕ 2L1 L2 10.20 Note that this equation specifies an upper limit on the value of the mutual inductance. We define the coefficient of coupling between the two inductors L1 and L2 as k = M 2L1 L2 10.21 and we note from Eq. (10.20) that its range of values is 0ⱕkⱕ1 10.22 This coefficient is an indication of how much flux in one coil is linked with the other coil; that is, if all the flux in one coil reaches the other coil, then we have 100% coupling and k = 1. For large values of k (i.e., k 7 0.5), the inductors are said to be tightly coupled, and for small values of k (i.e., k ⱕ 0.5), the coils are said to be loosely coupled. If there is no coupling, k = 0. The previous equations indicate that the value for the mutual inductance is confined to the range 0 ⱕ M ⱕ 2L1 L2 and that the upper limit is the geometric mean of the inductances L1 and L2 . 10.23 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 505 SECTION 10.2 The coupled circuit in Fig. 10.16a has a coefficient of coupling of 1 (i.e., k = 1). We wish to determine the energy stored in the mutually coupled inductors at time t = 5 ms. L1 = 2.653 mH and L2 = 10.61 mH. From the data the mutual inductance is E N E R G Y A N A LY S I S EXAMPLE 10.7 SOLUTION M = 2L1 L2 = 5.31 mH The frequency-domain equivalent circuit is shown in Fig. 10.16b, where the impedance values for XL1 , XL2 , and XM are 1, 4, and 2, respectively. The mesh equations for the network are then (2 + j1)I 1 - j2I 2 = 24 / 0° -j2I 1 + (4 + j4)I 2 = 0 Solving these equations for the two mesh currents yields I 1 = 9.41 / -11.31° A and I 2 = 3.33 / +33.69° A and therefore, i1(t) = 9.41 cos (377t - 11.31°) A i2(t) = 3.33 cos (377t + 33.69°) A At t = 5 ms, 377t = 1.885 radians or 108°, and therefore, i1(t = 5 ms) = 9.41 cos (108° - 11.31°) = -1.10 A i2(t = 5 ms) = 3.33 cos (108° + 33.69°) = -2.61 A Therefore, the energy stored in the coupled inductors at t = 5 ms is w(t)|t = 0.005 s = 1 1 (2.653)A10-3B(-1.10)2 + (10.61)A10-3B(-2.61)2 2 2 -(5.31)A10-3 B(-1.10)(-2.61) = (1.61)A10-3 B + (36.14)A10-3 B - (15.25)A10-3 B = 22.5 mJ M 2⍀ 24 cos 377t V ± – i1(t) L1 L2 i2(t) 4⍀ j4 ⍀ I2 4⍀ (a) j2 ⍀ 2⍀ 24 0° V ± – I1 j1 ⍀ (b) 505 Figure 10.16 Example of a magnetically coupled circuit drawn in the time and frequency domains. irwin10_491-540hr.qxd 506 28-07-2010 CHAPTER 10 12:10 Page 506 M A G N E T I C A L LY C O U P L E D N E T W O R K S Learning Assessment ANSWER: w(10 ms) = 39 mJ. E10.8 The network in Fig. E10.8 operates at 60 Hz. Compute the energy stored in the mutually coupled inductors at time t = 10 ms. 12 30° V ± – I1 –j2 ⍀ j1 ⍀ 2⍀ j2 ⍀ j2 ⍀ I2 2⍀ Figure E10.8 10.3 The Ideal Transformer Consider the situation illustrated in Fig. 10.17, showing two coils of wire wound around a single closed magnetic core. Assume a core flux , which links all the turns of both coils. In the ideal case we also neglect wire resistance. Let us now examine the coupling equations under the condition that the same flux goes through each winding and so, v1(t) = N1 d dt v2(t) = N2 d dt and and therefore, d N1 N1 dt v1 = = v2 N2 d N2 dt 10.24 Ampère’s law requires that C H ⴢ dl = ienclosed = N1 i1 + N2 i2 10.25 where H is the magnetic field intensity and the integral is over the closed path traveled by the flux around the transformer core. If H = 0, which is the case for an ideal magnetic core with infinite permeability, then N1 i1 + N2 i2 = 0 10.26 N2 i1 = i2 N1 10.27 or Note that if we divide Eq. (10.26) by N1 and multiply it by v1 , we obtain N2 v1 i2 = 0 N1 v1 i1 + Figure 10.17 Transformer employing a magnetic core. i1(t) v1(t) + – • A N1 N2 • i2(t) + v (t) – 2 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 507 SECTION 10.3 However, since v1 = AN1兾N2 Bv2 , THE IDEAL TRANSFORMER 507 v1 i1 + v2 i2 = 0 and hence the total power into the device is zero, which means that an ideal transformer is lossless. The symbol we employ for the ideal transformer is shown in Fig. 10.18a, and the corresponding equations are v1 N1 = v2 N2 10.28 N1 i1 + N2 i2 = 0 The normal power flow through a transformer occurs from an input current Ai1 B on the primary to an output current Ai2 B on the secondary. This situation is shown in Fig. 10.18b, and the corresponding equations are v1 N1 = v2 N2 10.29 N1 i1 = N2 i2 Note that although the voltage, current, and impedance levels change through a transformer, the power levels do not. The vertical lines between the coils, shown in the figures, represent the magnetic core. Although practical transformers do not use dots per se, they use markings specified by the National Electrical Manufacturers Association (NEMA) that are conceptually equivalent to the dots. Thus, our model for the ideal transformer is specified by the circuit in Fig. 10.18a and the corresponding Eq. (10.28), or alternatively by the circuit in Fig. 10.18b, together with Eq. (10.29). Therefore, it is important to note carefully that our model specifies the equations as well as the relationship among the voltages, currents, and the position of the dots. In other words, the equations are valid only for the corresponding circuit diagram. Thus, in a direct analogy to our discussion of the mutual inductance equations and their corresponding circuit, if we change the direction of the current or voltage or the position of the dots, we must make a corresponding change in the equations. The following material will clarify this critical issue. Consider now the circuit shown in Fig. 10.19. If we compare this circuit to that shown in Fig. 10.18b, we find that the direction of both the currents and voltages are the same. Hence the equations for the network are Figure 10.18 V1 N1 = V2 N2 Symbol for an ideal transformer: (a) primary and secondary currents into the dots; (b) primary current into, and secondary current out of, the dots. and I1 N2 = I2 N1 N1 : N2 + i1 N1 : N2 i2 v1 - + + v2 v1 - - i1 i2 + v2 - Ideal Ideal (a) (b) irwin10_491-540hr.qxd 508 28-07-2010 CHAPTER 10 12:10 Page 508 M A G N E T I C A L LY C O U P L E D N E T W O R K S Figure 10.19 N1 : N 2 Ideal transformer circuit used to illustrate input impedance. + + I1 V1 V2 - I2 ZL Ideal These equations can be written as V1 = N1 V2 N2 I1 = N2 I2 N1 10.30 Also note that ZL = V2 I2 and therefore the input impedance is Z1 = N1 2 V1 = a b ZL I1 N2 10.31 where ZL is reflected into the primary side by the turns ratio. If we now define the turns ratio as n = N2 N1 10.32 then the defining equations for the ideal transformer in this configuration are V1 = V2 n I 1 = nI 2 Z1 = 10.33 ZL n2 Care must be exercised in using these equations because the signs on the voltages and currents are dependent on the assigned references and their relationship to the dots. EXAMPLE Given the circuit shown in Fig. 10.20, we wish to determine all indicated voltages and currents. 10.8 I1 120 0° V –j4 ⍀ 18 ⍀ I2 4:1 ± – ± ± V1 V2 – – 2⍀ j1 ⍀ Figure 10.20 Ideal transformer circuit. SOLUTION Ideal Because of the relationships between the dots and the currents and voltages, the transformer equations are V1 = - V2 n and I 1 = -nI 2 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 509 SECTION 10.3 THE IDEAL TRANSFORMER 509 where n = 1兾4 . The reflected impedance at the input to the transformer is Z 1 = 42Z L = 16(2 + j1) = 32 + j16 ⍀ Therefore, the current in the source is I1 = 120 / 0° 18 - j4 + 32 + j16 = 2.33 / -13.5° A The voltage across the input to the transformer is then V1 = I 1 Z 1 = A2.33 / -13.5°B(32 + j16) = 83.49 / 13.07° V Hence, V2 is V2 = -nV1 1 A83.49 / 13.07°B 4 = 20.87 / 193.07° V = - The current I2 is I1 n = -4A2.33 / -13.5°B I2 = - = 9.33 / 166.50° A Learning Assessments E10.9 Compute the current I1 in the network in Fig. E10.9. –j2 ⍀ I1 2⍀ 12 0° V 2⍀ 1:2 + ± – –j2 ⍀ + V1 + 2⍀ V2 - Figure E10.9 ANSWER: I 1 = 3.07 / 39.81° A. Vo - - Ideal E10.10 Find Vo in the network in Fig. E10.9. ANSWER: Vo = 3.07 / 39.81° V. E10.11 Determine I1, I2, V1, and V2 in Fig. E10.11. ANSWER: I 1 = 2.12 / –25.6° A; I 2 = 8.49 / 154.42° A; V1 = 64.16 / 44.1° V; V2 = 16.04 / –135.9° V. j 16 ⍀ 32 ⍀ I1 4:1 I2 3⍀ j2 ⍀ 100 0° V + – + - V1 V2 – + 20 –20° V Figure E10.11 –j 24 ⍀ Ideal + – irwin10_491-540hr.qxd 510 28-07-2010 CHAPTER 10 12:10 Page 510 M A G N E T I C A L LY C O U P L E D N E T W O R K S E10.12 Determine Vo in Fig. E10.12. 32 ⍀ j 16 ⍀ ANSWER: Vo = 24.95 / –62.65° V. 3⍀ 4:1 + j2 ⍀ 100 0° V + – Vo 20 –20° V + – – Figure E10.12 Ideal –j 24 ⍀ E10.13 Determine Vo in Fig. E10.13. 2⍀ j4 ⍀ ANSWER: Vo = 93.68 / –83° V. 1:2 –j 8 ⍀ 10 ⍀ 4:1 0.5 ⍀ + 50 0° V + – + – Vo 10 30° V – Ideal j 10 ⍀ 6⍀ Ideal –j 0.5 ⍀ Figure E10.13 E10.14 If Vo = 10 / 30° V in Fig. E10.14, find VS. 2⍀ 4⍀ 2:1 2⍀ ANSWER: VS = 32.34 / –125.3° V. j4 ⍀ + VS + – j5 ⍀ –j 5 ⍀ 10 ⍀ Vo – Ideal Figure E10.14 Another technique for simplifying the analysis of circuits containing an ideal transformer involves the use of either Thévenin’s or Norton’s theorem to obtain an equivalent circuit that replaces the transformer and either the primary or secondary circuit. This technique usually requires more effort, however, than the approach presented thus far. Let us demonstrate this approach by employing Thévenin’s theorem to derive an equivalent circuit for the transformer and primary circuit of the network shown in Fig. 10.21a. The equations for the transformer in view of the direction of the currents and voltages and the position of the dots are I 1 = nI 2 V1 = V2 n irwin10_491-540hr.qxd 28-07-2010 12:10 Page 511 SECTION 10.3 Z1 V S1 ± – I1 1 1:n + + V1 V2 - Z2 I2 Z1 ± VS2 – VS1 ± – I1 - I2 + V1 - 1' Ideal 2' Ideal (a) (b) n2Z1 nVS1 ± – 2 THE IDEAL TRANSFORMER 2 Z2 I1 V2=nV1 I2=––– n Z1 ± VS2 – 2' VS1 ± – 511 2 + + V2 Voc - 2' Z2/n2 1 V2 V1=––– n I1=nI2 S2 ± V – ––– n 1' (d) (c) Forming a Thévenin equivalent at the secondary terminals 2 = 2¿, as shown in Fig. 10.21b, we note that I2 = 0 and therefore I1 = 0. Hence Voc = V2 = nV1 = nVS1 The Thévenin equivalent impedance obtained by looking into the open-circuit terminals with VS1 replaced by a short circuit is Z1 , which when reflected into the secondary by the turns ratio is Figure 10.21 Circuit containing an ideal transformer and some of its equivalent networks. Z Th = n2 Z 1 Therefore, one of the resulting equivalent circuits for the network in Fig. 10.21a is as shown in Fig. 10.21c. In a similar manner, we can show that replacing the transformer and its secondary circuit by an equivalent circuit results in the network shown in Fig. 10.21d. It can be shown in general that when developing an equivalent circuit for the transformer and its primary circuit, each primary voltage is multiplied by n, each primary current is divided by n, and each primary impedance is multiplied by n2 . Similarly, when developing an equivalent circuit for the transformer and its secondary circuit, each secondary voltage is divided by n, each secondary current is multiplied by n, and each secondary impedance is divided by n2 . Powers are the same, whether calculated on the primary or secondary side. Recall from our previous analysis that if either dot on the transformer is reversed, then n is replaced by -n in the equivalent circuits. In addition, note that the development of these equivalent circuits is predicated on the assumption that removing the transformer will divide the network into two parts; that is, there are no connections between the primary and secondary other than through the transformer. If any external connections exist, the equivalent circuit technique cannot in general be used. Finally, we point out that if the primary or secondary circuits are more complicated than those shown in Fig. 10.21a, Thévenin’s theorem may be applied to reduce the network to that shown in Fig. 10.21a. Also, we can simply reflect the complicated circuit component by component from one side of the transformer to the other. Given the circuit in Fig. 10.22a, we wish to draw the two networks obtained by replacing the transformer and the primary, and the transformer and the secondary, with equivalent circuits. Due to the relationship between the assigned currents and voltages and the location of the dots, the network containing an equivalent circuit for the primary and the network containing an equivalent circuit for the secondary are shown in Figs. 10.22b and c, respectively. The reader should note carefully the polarity of the voltage sources in the equivalent networks. EXAMPLE SOLUTION 10.9 irwin10_491-540hr2.qxd 512 13-09-2010 CHAPTER 10 13:00 Page 512 M A G N E T I C A L LY C O U P L E D N E T W O R K S Figure 10.22 –j3 ⍀ 4⍀ Example circuit and two equivalent circuits. 1 + 1:2 ± – 12 0° V I1 12 ⍀ 2 V1 j4 ⍀ + - – ± I2 V2 48 30° V 1' 2' Ideal (a) 16 ⍀ 24 0° V –j12 ⍀ – ± 2 12 ⍀ j4 ⍀ – ± I2 V2 48 30° V 2' (b) 4⍀ 12 0° V ± – –j3 ⍀ 1 V1 3⍀ j1 ⍀ I1 ± – 24 30° V 1' (c) Problem-Solving Strategy Circuits Containing Ideal Transformers Step 1. Carefully examine the circuit diagram to determine the assigned voltage polarities and current directions in relation to the transformer dots. ● If both voltages are referenced positive at the dotted terminals or undotted terminals, then v1兾v2 = N1兾N2 . If this is not true, then v1兾v2 = -N1兾N2 . If one current is defined as entering a dotted terminal and the other current is defined as leaving a dotted terminal, then N1 i1 = N2 i2 . If this condition is not satisfied, then N1 i1 = -N2 i2 . Step 2. If there are no electrical connections between two transformer windings, reflect all circuit elements on one side of the transformer through to the other side, thus eliminating the ideal transformer. Be careful to apply the statements above when reflecting elements through the transformer. Remember that impedances are scaled in magnitude only. Apply circuit analysis techniques to the circuit that results from eliminating all ideal transformers. After this circuit has been analyzed, reflect voltages and currents back through the appropriate ideal transformers to find the answer. Step 3. As an alternative approach, use Thévenin’s or Norton’s theorem to simplify the circuit. Typically, calculation of the equivalent circuit eliminates the ideal transformer. Solve the simplified circuit. ● irwin10_491-540hr.qxd 28-07-2010 12:10 Page 513 SECTION 10.3 THE IDEAL TRANSFORMER 513 Step 4. If there are electrical connections between two transformer windings, use nodal analysis or mesh analysis to write equations for the circuits. Solve the equations using the proper relationships between the voltages and currents for the ideal transformer. EXAMPLE Let us determine the output voltage Vo in the circuit in Fig. 10.23a. 10.10 SOLUTION We begin our attack by forming a Thévenin equivalent for the primary circuit. From Fig. 10.23b we can show that the open-circuit voltage is Voc = 24 / 0° (-j4) - 4 / -90° 4 - j4 = 12 - j8 = 14.42 / -33.69° V The Thévenin equivalent impedance looking into the open-circuit terminals with the voltage sources replaced by short circuits is Z Th = (4)(-j4) 4 - j4 + 2 Figure 10.23 = 4 - j2 ⍀ Example network and other circuits used to derive an equivalent network. The circuit in Fig. 10.23a thus reduces to that shown in Fig. 10.23c. Forming an equivalent circuit for the transformer and primary results in the network shown in Fig. 10.23d. 4 –90° V 4⍀ 2⍀ + 24 0° V ± – –j4 ⍀ 2⍀ 1:2 ±– j3 ⍀ + V1 + 2⍀ V2 - - Vo - Ideal (a) 4⍀ 4 –90° V 2⍀ 4⍀ –j2 ⍀ + 24 0° V ± – – j4 ⍀ Voc + ± – 14.42 –33.69° V - 2⍀ 1:2 ±– + V1 + V2 - - (c) 16 ⍀ –j8 ⍀ 2⍀ j3 ⍀ + + 2⍀ 28.84 –33.69° V Vo - (d) 2⍀ Vo - Ideal (b) j3 ⍀ irwin10_491-540hr.qxd 514 28-07-2010 CHAPTER 10 12:10 Page 514 M A G N E T I C A L LY C O U P L E D N E T W O R K S Therefore, the voltage Vo is Vo = -28.84 / -33.69° 20 - j5 (2) = 2.80 / 160.35° V Learning Assessments E10.15 Given the network in Fig. E10.15, form an equivalent circuit for the transformer and secondary, and use the resultant network to compute I1 . I1 36 0° V 2⍀ – j2 ⍀ 2⍀ 1:2 ± – ± – Figure E10.15 ANSWER: I 1 = 13.12 / 38.66° A. 12 0° V Ideal E10.16 Given the network in Fig. E10.16, form an equivalent circuit for the transformer and primary, and use the resultant network to find Vo . 2⍀ ANSWER: Vo = 3.12 / 38.66° V. 4 0° V – j2 ⍀ 1:2 ±– + 12 0° V ± – 2⍀ Vo - Figure E10.16 EXAMPLE Ideal Determine I1 , I2 , V1 , and V2 in the network in Fig. 10.24. 10.11 SOLUTION The nodal equations at nodes 1 and 2 are V1 - V2 10 - V1 = + I1 2 2 V1 - V2 V2 I2 + = 2 2j The transformer relationships are V2 = 2V1 and I1 = 2I2 . The first nodal equation yields I1 = 5 A and therefore I2 = 2.5 A. The second nodal equation, together with the constraint equations specified by the transformer, yields V1 = 15 / 63° V and V2 = 2 15 / 63° V. 2⍀ Figure 10.24 Circuit used in Example 10.11. 2⍀ I1 1:2 I2 1 10 0° V ± – 2 + + V1 - V2 Ideal j2 ⍀ irwin10_491-540hr.qxd 28-07-2010 12:10 Page 515 SECTION 10.4 S A F E T Y C O N S I D E R AT I O N S 515 Learning Assessment E10.17 Determine I1 , I2 , V1 , and V2 in the network in Fig. E10.17. ANSWER: I 1 = 3.08 / -13.7° A; 2⍀ 2⍀ I1 1:2 Figure E10.17 ± – 2⍀ V1 = 0.85 / 20° V; V2 = 1.71 / -160° V. + + 10 0° V I 2 = 1.54 / 166.3° A; I2 V1 - V2 - j2 ⍀ Ideal Before we move on to the next topic, let’s return to Faraday’s law. For the ideal d d transformer, Faraday’s law tells us that v1(t) = N1 and v2(t) = N2 . What if a dc dt dt voltage is applied to our transformer? In that case, the magnetic flux is a constant, v1 = v2 = 0, and our transformer is not very useful. What if an ac voltage is applied to our transformer? The magnetic flux is sinusoidal and time-varying. Transformers allow the ac voltage value to be stepped up or down easily and efficiently; it is much more difficult to efficiently step up or down the dc voltage value. The ease with which transformers allow us to change the voltage level is one of the main reasons that ac voltages and currents are utilized to transmit the bulk of the world’s electrical power. Transistors are used extensively in modern electronic equipment to provide a low-voltage power supply. As examples, a common voltage level in computer systems is 5 V dc, portable radios use 9 V dc, and military and airplane equipment operates at 28 V dc. When transformers are used to connect these low-voltage transistor circuits to the power line, there is generally less danger of shock within the system because the transformer provides electrical isolation from the line voltage. However, from a safety standpoint, a transformer, although helpful in many situations, is not an absolute solution. When working with any electrical equipment, we must always be vigilant to minimize the dangers of electrical shock. In power electronics equipment or power systems, the danger is severe. The problem in these cases is that of high voltage from a low-impedance source, and we must constantly remember that the line voltage in our homes can be lethal. Consider now the following example, which illustrates a hidden danger that could surprise even the experienced professional, with devastating consequences. Two adjacent homes, A and B, are fed from different transformers, as shown in Fig. 10.25a. A surge on the line feeding house B has caused the circuit breaker X-Y to open. House B is now left without power. In an attempt to help his neighbor, the resident of house A volunteers to connect a long extension cord between a wall plug in house A and a wall plug in house B, as shown in Fig. 10.25b. Later, the line technician from the utility company comes to reconnect the circuit breaker. Is the line technician in any danger in this situation? 10.4 Safety Considerations EXAMPLE 10.12 irwin10_491-540hr.qxd 516 28-07-2010 CHAPTER 10 12:10 Page 516 M A G N E T I C A L LY C O U P L E D N E T W O R K S SOLUTION Unaware of the extension cord connection, the line technician believes that there is no voltage between points X and Z. However, because of the electrical connection between the two homes, 7200 V rms exists between the two points, and the line technician could be seriously injured or even killed if he comes in contact with this high voltage. Figure 10.25 Diagrams used in Example 10.12 (voltages in rms). X Y 120 V 7200 V 0V 0V 7200 V Z A B (a) 120 V X 120 V 7200 V 7200 V Y 7200 V Z A B (b) 10.5 The following examples demonstrate several applications for transformers. Application Examples APPLICATION EXAMPLE 10.13 SOLUTION Consider the problem of transporting 24 MW over a distance of 100 miles (160.9 km) using a two-conductor line. Determine the requisite conductor radius to achieve a transmission efficiency of 95%, considering only the line resistance, if the line operates at (a) 240 V rms or (b) 240 kV rms. Assume that conductor resistivity is = 8 * 10-8 ⍀ -m. a. At 240 V: I = P 24 M = = 100 kA rms V 240 If = 95%, Ploss = 0.05(24 M) = 1.2 MW = I 2R Therefore, R = Ploss I2 = 1.2 M (100k)2 = 1.2 * 10-4⍀ irwin10_491-540hr.qxd 28-07-2010 12:10 Page 517 SECTION 10.5 A P P L I C AT I O N E X A M P L E S 517 Since R = (8 * 10-8)(2 * 160.9 * 103) l = A A A = 0.25744 = 214.5 m2 = r2 1.2 * 10-4 Therefore, r = 8.624 m (which is a huge conductor and totally impractical!) b. At 240 kV rms: I = 100 A rms and R = A = 1.2 * 106 (100)2 = 120 ⍀ 0.25744 = 2.145 * 10-4 m 120 and r = 0.8264 cm (which is a very practical value!) The point of this example is that practical transmission of bulk electrical energy requires operation at high voltage. What is needed is an economical device that can efficiently convert one voltage level to another. Such a device is, as we have shown, the power transformer. The local transformer in Fig. 10.26 provides the last voltage stepdown in a power distribution system. A common sight on utility poles in residential areas, it is a single-phase transformer that typically has a 13.8-kV rms line to neutral on its primary coil, and a center tap secondary coil provides both 120 V rms and 240 V rms to service several residences. A typical example of this transformer, often referred to as a “pole pig,” is shown in Fig. 10.27. a 1:n Substation Figure 10.26 + + 120 V rms 240 V rms + 120 V rms - + 13.8 kV rms - APPLICATION EXAMPLE 10.14 n Local step-down transformer Let us find the turns ratio necessary to produce the 240-V rms secondary voltage. Assuming that the transformer provides 200-A rms service to each of 10 houses, let us determine the minimum power rating for the transformer and the maximum current in the primary. Local transformer subcircuit with center tap. irwin10_491-540hr.qxd 518 28-07-2010 CHAPTER 10 12:10 Page 518 M A G N E T I C A L LY C O U P L E D N E T W O R K S SOLUTION The turns ratio is given by n = V2 240 1 = = V1 13,800 57.5 If IH is the maximum current per household, then the maximum primary current is I1 = nI2 = nA10IH B = 34.78 A rms The maximum power delivered to the primary is then S1 = V1 I1 = (13,800)(34.78) = 480 kVA. Therefore, the transformer must have a power rating of at least 480 kVA. Figure 10.27 A residential utility transformer. (tomba/iStockphoto) APPLICATION EXAMPLE 10.15 Your electric toothbrush sits innocuously in its cradle overnight. Even though there are no direct electrical connections between the cradle and the toothbrush, the internal batteries are being recharged. How can this be? SOLUTION Mutually coupled inductors is the answer! One coil is in the cradle and energized by an ac source. The second coil is in the bottom of the toothbrush itself. When the toothbrush is mounted in the cradle, the two coils are physically close and thus mutually coupled, as shown in Fig. 10.28. Let’s investigate a reasonable design for these coils. First, we assume that the coils are poorly coupled with a coupling coefficient of k = 0.25. Second, the coil in the cradle is driven at 120 / 0° V rms. Third, the coil in the toothbrush should generate 6 / 0° V at 100 mA rms in order to charge the battery. To keep the power “relatively” low, we will limit the primary current to only 0.5 A rms. Finally, to simplify our analysis, we will assume that I 1 and I 2 are in phase. First we develop loop equations for our circuit, which are V1 = jL1 I 1 - jMI 2 V2 = jMI 1 - jL2 I 2 10.34 where V1 = 120 / 0° V rms and V2 = 6 / 0° V rms. By defining a new variable ␣ such that L2 = ␣2L1 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 519 SECTION 10.5 A P P L I C AT I O N E X A M P L E S 519 Toothbrush Coil 2 jM I1 I2 Coil 1 + V1 120 0° V rms Cradle ± – jL1 jL2 V2 Battery charger - (a) (b) we can eliminate L2 in Eq. (10.34). Hence, Figure 10.28 V1 = jL1 I 1 - jk␣L1 I 2 10.35 V2 = jk␣L1 I 1 - j␣2L1 I 2 Taking the ratio of each side of Eq. (10.35), we can eliminate and L1: V1 I 1 - k␣I 2 120 = = 20 = V2 6 k␣I 1 - ␣2I 2 If we now substitute the design parameter values listed above for I 1 , I 2 and k and then solve for ␣, we find 20I2 ␣2 - A20kI1 + kI2 B␣ + I1 = 0 which yields ␣ = b 0.246 1.02 (We have used our restriction that I 1 and I 2 are in phase to convert current phasors to magnitudes.) Choosing the smaller value for ␣ is the same as choosing a smaller L2 . Hence, this is the result we select since the resulting coil will require fewer turns of wire, reducing cost, weight, and size. Next, using Eq. (10.35), we can solve for the product L1: V1 = L1(0.5) - (0.25)(0.246)L1(0.1) = 0.494L1 To investigate the effect of on L1 , we use the value for V1 and the relationship between the two inductors, for the given ␣. In Table 10.1, L1 and L2 have been calculated for a collection of values. Note that a 60-Hz excitation requires huge values for the inductances TABLE 10.1 A listing of frequency choices and the resulting inductances. FREQUENCY (Hz) FREQUENCY (rad/s) L1 L2 60 20k 100k 377 126k 628k 693 mH 2.01 mH 416 H 39.0 mH 117 H 23.4 H The electric toothbrush: (a) a nonartist’s conceptual drawing and (b) a circuit schematic. irwin10_491-540hr.qxd 520 28-07-2010 CHAPTER 10 12:10 Page 520 M A G N E T I C A L LY C O U P L E D N E T W O R K S that are completely impractical. Therefore, the table skips past the entire audible range (there’s no reason to have to listen to your toothbrush recharging) to 20 kHz. Here the inductance values are much more reasonable but still considerable. However, at 100 kHz, the total inductance is just a few hundred microhenrys. These are very reasonable values and ones which we will use. The final question is this: if we have a 60-Hz sinusoid at the wall outlet, how do we obtain 100 kHz? We add a voltage-controlled switch as shown in Fig. 10.29 that is turned on and off at a rate of 100 kHz. The result is a pulsing voltage applied to the inductor at 100 kHz. Although the result is not exactly a 100-kHz sine wave, it is effective. Figure 10.29 M i1(t) A switch, turned on and off at a 100-kHz rate, can emulate a high-frequency ac input for our toothbrush application. i2(t) + L1 170 cos 377t V v2(t) L2 Battery charger - ± – 100 kHz APPLICATION EXAMPLE 10.16 As shown in Fig. 10.30, two circuits are placed in close proximity: a high-current ac circuit and a low-current dc circuit. Since each circuit constitutes a loop, we should expect a little inductance in each circuit. Because of their proximity, we could also anticipate some coupling. In this particular situation the inductance in each loop is 10 nH, and the coupling coefficient is k = 0.1. Let us consider two scenarios. In the first case, the ac circuit contains an ac motor operating at 60 Hz. In the second case, the ac circuit models a FM radio transmitter operating at 100 MHz. We wish to determine the induced noise in the dc circuit for both cases. Which scenario produces the worst inductively coupled noise? Why? Figure 10.30 A circuit model for an ac and a dc circuit that are physically close enough to have mutual coupling. 5 0° A 340 0° V jLAC ac LOAD ± – LDC 5 Vdc ± – k dc LOAD irwin10_491-540hr.qxd 28-07-2010 12:10 Page 521 SECTION 10.6 DESIGN EXAMPLES 521 SOLUTION The voltage induced into the dc circuit is noise and is known to be Vnoise = jMI AC = jk1LAC LDCI AC We are concerned only with the magnitude of the noise. Given the model parameters listed above, the noise voltage magnitude is Vnoise = 2f(0.1)A10-8 B(5) = 3.14 * 10-8f V For the ac motor scenario, f = 60 Hz and the noise voltage is 1.88 V—essentially zero when compared against the 5-V dc input. However, when modeling a FM radio transmitter operating at 100 MHz the noise voltage is 3.14 V. That’s more than 60% of the 5-V dc level! Thus, we find that magnetically-induced noise is much worse for high-frequency situations. It should be no surprise then that great care is taken to magnetically “shield” high-frequency–high-current circuitry. 10.6 Design Examples A linear variable differential transformer (LVDT), is commonly used to measure linear movement. LVDTs are useful in a wide range of applications such as measuring the thickness of thin material sheets and measuring the physical deformation of objects under mechanical load. (A Web search on LVDT will yield a multitude of other example applications with explanations and photographs.) As shown in Figs. 10.31a and b, the LVDT is just a coupled inductor apparatus with one primary winding and two secondaries that are wound and connected such that their induced voltages subtract. All three windings are contained in a hollow cylinder that receives a rod, usually made of steel or iron, that is physically attached to whatever it is that’s moving. The presence of the rod drastically increases the coupling coefficient between the windings. Let us investigate how the LVDT output voltage is related to displacement and how the LVDT is driven. Then, we will design our own LVDT, driven at 10 V rms, 2 kHz, such that at 100% travel, the output voltage magnitude equals that of the input voltage. M12 IP IS Figure 10.31 + LS Magnetic rod Secondary coil Primary coil Secondary coil (a) Vin ± – LP Vo LS - M13 (b) DESIGN EXAMPLE 10.17 Two representations of the standard LVDT: (a) the cutaway view and (b) the circuit diagram. irwin10_491-540hr.qxd 522 28-07-2010 CHAPTER 10 12:10 Page 522 M A G N E T I C A L LY C O U P L E D N E T W O R K S SOLUTION Typically, the primary winding of the LVDT is excited by an ac sinusoid in the range of 3 to 30 V rms at frequencies between 400 and 5000 Hz. Since we only need to measure the output voltage directly with a voltmeter, no external load is necessary. The null position for the rod is dead center between the secondary windings. In that position the coupling between the primary and each secondary is identical, and the output voltage is zero. Should the rod move in either direction, the coupling will change linearly, as will the output voltage magnitude. The direction of travel is indicated by the relative phase of the output. Our LVDT design begins with the circuit in Fig. 10.31b where the mutual coupling coefficient for each secondary winding varies as shown in Fig. 10.32. To create a linear relationship between displacement and output voltage, we restrict the nominal travel to that portion of Fig. 10.32 where the coupling coefficient is linear with displacement. Therefore, in this design, 100% travel will correspond to a coupling coefficient of 0.8. Applying KVL to the primary loop yields Vin = jLP I P + jM13 I S - jM12 I S 10.36 At the secondary, the KVL equation is 2AjLS BI S + jM13 I P - jM12 I P + Vo = 0 10.37 With no load at the output, I S = 0 and Eqs. (10.36) and (10.37) reduce to Vin = jLP I P and Vo = I P j CM12 - M13 D 10.38 Solving these for the output voltage and recognizing that MlX = klX CLP LS D ,we obtain 0.5 Vo = Vin LS C k - k13 D B LP 12 10.39 We can express the coupling coefficients for each secondary in terms of the percent of travel: k12 = b for 0 6 x 6 100 for x 6 0 k13 = b 0.008x 0 for -100 6 x 6 0 for x 7 0 1.2 Figure 10.32 [k12-k13] 1 0.8 Coupling Coefficient Coupling coefficients for each secondary winding and the coupling difference. It is the difference that will determine the output voltage magnitude. 0.008x 0 k12 k13 0.6 0.4 0.2 0 –0.2 –200 –150 –100 –50 0 Travel 50 100 150 200 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 523 SECTION 10.6 DESIGN EXAMPLES And, finally, assuming that the input voltage has zero phase angle, the output voltage can be expressed as Vo = Vin LS LS [0.008x] = Vin [0.008x] / 0° B LP B LP 0 6 x 6 100 LS LS Vo = Vin [-0.008x] = Vin [+0.008x] / -180° B LP B LP 10.40 -100 6 x 6 0 Note the phase angle difference for positive versus negative travel. To complete our analysis, we must determine a value for the secondary to primary inductance ratio. At 100% travel, the magnitudes of the input and output voltages are equal and k = 0.8. Using this information in Eq. (10.40), we find that the inductor ratio must be LS兾LP = 1.252 = 1.5625. To determine actual values for the inductances, we must consider the input current we will tolerate at the primary. We would prefer a relatively small current, because a large current would require large-diameter wire in the primary winding. Let us choose a primary current of 25 mA rms with an excitation of 10 V rms at 2000 Hz. From Eq. (10.38), the primary inductance will be LP = Vin 10 = 31.8 mH = IP 2(2000)(0.025) which yields a secondary inductance of LS = 1.5625LP = 49.7 mH The selection of the two inductances completes this design. The next example illustrates a technique for employing a transformer in a configuration that will extend the life of a set of Christmas tree lights. The bulbs in a set of Christmas tree lights normally operate at 120 V rms. However, they last much longer if they are instead connected to 108 V rms. Using a 120 V-12 V transformer, let us design an autotransformer that will provide 108 V rms to the bulbs. DESIGN EXAMPLE 10.18 The two-winding transformers we have presented thus far provide electrical isolation SOLUTION between primary and secondary windings, as shown in Fig. 10.33a. It is possible, however, to interconnect primary and secondary windings serially, creating a three-terminal device, known as an autotransformer, as shown in Fig. 10.33b and represented in Fig. 10.33c. As we shall see, this arrangement offers certain practical advantages over the isolated case. Note that the three-terminal arrangement is essentially one continuous winding with an internal tap. To reduce the voltage from 120 V rms to 108 V rms, the two coils must be connected such that their voltages are in opposition to each other, corresponding to a subtractive connection (in Fig. 10.33b), as shown in Fig. 10.34. In this arrangement, the voltage across both coils is Vo = V1 - V2 = 120 - 12 = 108 V rms and the lights are simply connected across both coils. 523 irwin10_491-540hr.qxd 524 28-07-2010 CHAPTER 10 12:10 Page 524 M A G N E T I C A L LY C O U P L E D N E T W O R K S X N1 X Y N1 N1 Y N2 Z N2 Z N2 Additive connection (a) Subtractive connection (b) X X N1 N1 Y Y N2 N2 Z Z Additive connection Subtractive connection (c) Figure 10.33 Autotransformer: (a) normal two-winding transformer with adjacent windings; (b) two-winding transformer interconnected to create a single-winding, three-terminal autotransformer; (c) symbolic representation of (b). Figure 10.34 - Autotransformer for low-voltage Christmas tree lights. 120 V rms ± – V2 12 V rms + + Vo=108 V rms V1 120 V rms - DESIGN EXAMPLE 10.19 + Christmas lights - Many electronics products today are powered by low-power ac to dc converters. (These units simply convert an ac signal at the input to a constant dc signal at the output.) They are normally called wall transformers and plug directly into a 120 V rms utility outlet. They typically have dc output voltages in the range of 5 to 18 V. As shown in Fig. 10.35, there are three basic components in a wall transformer: a simple transformer, an ac to dc converter, and a controller. A particular wall transformer is required that has a dc output of 9 V and a maximum power output of 2 W at an efficiency of only 60%. In addition, the ac to dc converter requires a peak ac voltage input of 12 V for proper operation. We wish to design the transformer by selecting its turns ratio and current rating. irwin10_491-540hr.qxd 28-07-2010 12:10 Page 525 SUMMARY Controller Figure 10.35 ac to dc Converter A block diagram for a simple wall transformer. These devices convert ac voltages (typically 120 V rms) to a dc voltage at a fairly low power level. + V1 120 0° V rms ± – V2 525 + - VDC 9V - First we consider the necessary turns ratio for the transformer. We must determine the volt- SOLUTION age ratio, V2兾V1 . From the specifications, V2 must have a peak value of at least 12 V. We will include some safety margin and design for V2 around 13.5 V. Since V1 is 120 V rms, its peak value is 169.7 V. Therefore, V2 169.7 = = 12.6 V1 13.5 n = Thus, the V2兾V1 ratio is 12.6. We will use a turns ratio of 12.5:1, or 25:2. Next we consider the power requirement. The maximum load is 2 W. At an efficiency of 60%, the maximum input power to the unit is Pin = Pout 2 = = 3.33 W 0.6 At 120 V rms, the input current is only Iin = Pin 3.33 = = 27.8 mA rms Vin 120 Therefore, specifying a transformer with a turns ratio of 25:2 and a current rating of 100 mA rms should provide an excellent safety margin. • SUMMARY ■ Mutual inductance Mutual inductance occurs when inductors are placed in close proximity to one another and share a common magnetic flux. ■ The dot convention for mutual inductance The dot convention governs the sign of the induced voltage in one coil based on the current direction in another. ■ The relationship between the mutual inductance and self-inductance of two coils An energy analysis indicates that M = k1L1 L2 , where k, the coefficient of coupling, has a value between 0 and 1. ■ The ideal transformer An ideal transformer has infinite core permeability and winding conductance. The voltage and current can be transformed between the primary and secondary ends based on the ratio of the number of winding turns between the primary and secondary. ■ The dot convention for an ideal transformer The dot convention for ideal transformers, like that for mutual inductance, specifies the manner in which a current in one winding induces a voltage in another winding. ■ Equivalent circuits involving ideal transformers Based on the location of the circuits’ unknowns, either the primary or secondary can be reflected to the other side of the transformer to form a single circuit containing the desired unknown. The reflected voltages, currents, and impedances are a function of the dot convention and turns ratio. irwin10_491-540hr.qxd 526 • 28-07-2010 CHAPTER 10 12:10 Page 526 M A G N E T I C A L LY C O U P L E D N E T W O R K S PROBLEMS 10.4 Given the network in Fig. P10.4, 10.1 Given the network in Fig. P10.1, (a) write the equations for va(t) and vb(t). (a) find the equations for va(t) and vb(t). (b) write the equations for vc(t) and vd(t). (b) find the equations for vc(t) and vd(t). i1(t) + va(t) - M i2(t) - - + - L2 vd(t) vb(t) vc(t) L1 + + M i1(t) Figure P10.1 + va(t) - i2(t) + L2 vb(t) vd(t) vc(t) L1 + - - - + Figure P10.4 10.5 Find Vo in the network in Fig. P10.5. j1 ⍀ 1⍀ 1⍀ + 10.2 Given the network in Fig. P10.2, (a) find the equations for va(t) and vb(t). 12 0° V ± – j1 ⍀ j2 ⍀ 1⍀ Vo (b) find the equations for vc(t) and vd(t). i1(t) + va(t) - M - i2(t) - - Figure P10.5 + L2 vd(t) vb(t) vc(t) L1 + + - 10.6 Find Vo in the network in Fig. P10.6. j1 ⍀ 1⍀ Figure P10.2 1⍀ –j1 ⍀ + 12 0° V ± – j2 ⍀ j1 ⍀ 2⍀ Vo - Figure P10.6 10.3 Given the network in Fig. P10.3, (a) write the equations for va(t) and vb(t). (b) write the equations for vc(t) and vd(t). i1(t) + va(t) - Figure P10.3 - vc(t) L1 + M 10.7 Find Vo in the circuit in Fig. P10.7. i2(t) - 1⍀ + L2 vd(t) vb(t) + j1 ⍀ 2⍀ + 10 0° V ± – - j2 ⍀ j2 ⍀ 1⍀ Vo - Figure P10.7 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 527 PROBLEMS 10.8 Determine Vo in the network in Fig. P10.8. –j 2 ⍀ 3⍀ j1 ⍀ j4 ⍀ j4 ⍀ + 24 0° V + – 6⍀ 1⍀ 2⍀ 1⍀ Vo – Figure P10.8 10.9 Find Vo in the circuit in Fig. P10.9. j1 ⍀ 2⍀ j1 ⍀ + 10 0° V ± – j2 ⍀ 1⍀ j2 ⍀ 1⍀ Vo - Figure P10.9 10.10 Find Vo in the network in Fig. P10.10. – j1 ⍀ j2 ⍀ 2⍀ –j1 ⍀ 1⍀ + 24 0° V ± – 2⍀ j4 ⍀ j6 ⍀ 1⍀ Vo - Figure P10.10 10.11 Find Vo in the circuit in Fig. P10.11. j2 ⍀ 2⍀ j4 ⍀ j8 ⍀ + 24 0° V ± – – j2 ⍀ 6⍀ 8⍀ 4⍀ Vo - Figure P10.11 527 irwin10_491-540hr.qxd 528 28-07-2010 CHAPTER 10 12:10 Page 528 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10.12 Find I o in the circuit in Fig. P10.12. –j 1 ⍀ –j 2 ⍀ 12 ⍀ –+ 12 0° V 4⍀ –j 1 ⍀ j3⍀ 4⍀ – + 6⍀ 6 0° V Io Figure P10.12 10.13 Find Vo in the network in Fig. P10.13. j1 ⍀ –j1 ⍀ 1⍀ + ± – 10 0° V j2 ⍀ 1⍀ j1 ⍀ Vo 2⍀ - Figure P10.13 10.14 Find Vo in the circuit in Fig. P10.14. 2⍀ 12 0° V – + 4⍀ –j 1 ⍀ –j 2 ⍀ –j 1 ⍀ 12 ⍀ 24 0° V + + – 4⍀ j4⍀ j3⍀ 2⍀ Vo – Figure P10.14 10.15 Find Vo in the network in Fig. P10.15. 6⍀ 24 0° V – + 6⍀ j4⍀ j2⍀ j4⍀ 6⍀ + 4⍀ Figure P10.15 2 0° V –j 1 ⍀ Vo – 2⍀ 2⍀ 3 0° V irwin10_491-540hr.qxd 28-07-2010 12:10 Page 529 PROBLEMS 10.16 Find Vo in the network in Fig. P10.16. j1 ⍀ + j2 ⍀ 1⍀ 10 0° A j2 ⍀ 2⍀ Vo - j1 ⍀ 10 30° A Figure P10.16 10.17 Find Vo in the network in Fig. P10.17. 4 0° V 2 0° V j1⍀ –j 1 ⍀ + 3⍀ 24 0° V + – Vo 1⍀ 6⍀ j2⍀ – 3⍀ –j 4 ⍀ j2⍀ 1⍀ Figure P10.17 10.18 Find Vo in the network in Fig. P10.18. 2⍀ –j 1 ⍀ j2⍀ 3⍀ 12 0° V + – j2⍀ j1⍀ + 2 0° V Vo – Figure P10.18 529 irwin10_491-540hr.qxd 530 28-07-2010 CHAPTER 10 12:10 Page 530 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10.19 Find I o in the circuit in Fig. P10.19. j1 ⍀ –j4 ⍀ 2⍀ –j2 ⍀ j2 ⍀ 32 0° V 10.23 Find Vo in the network in Fig. P10.23. j2 ⍀ –j2 ⍀ 3⍀ + 2⍀ j2 ⍀ j2 ⍀ j2 ⍀ –j1 ⍀ Io ± – Vo 4 0° A - Figure P10.23 2⍀ 10.24 Find Vo in the network in Fig. P10.24. Figure P10.19 j1 ⍀ j2 ⍀ 1⍀ j2 ⍀ + 10.20 Find Vo in the network in Fig. P10.20. j1 ⍀ ± – –j1 ⍀ –j1 ⍀ 24 0° V 1⍀ 1⍀ 1⍀ j1 ⍀ j1 ⍀ 4 0° A Vo - + j1 ⍀ Figure P10.24 1⍀ Vo 1⍀ 10.25 Find Vo in the network in Fig. P10.25. j1 ⍀ 6 0° A 2⍀ j2 ⍀ 1⍀ j2 ⍀ - + –j2 ⍀ ± – Figure P10.20 –j1 ⍀ 24 0° V 1⍀ Vo - 10.21 Find I o in the circuit in Fig. P10.21. Figure P10.25 j1 ⍀ –j1 ⍀ j2 ⍀ j2 ⍀ 1⍀ 4 0° A 10.26 Find Vo in the network in Fig. P10.26. j1 ⍀ j1 ⍀ 2⍀ – j1 ⍀ 1⍀ + 1⍀ –j2 ⍀ –j1 ⍀ Io j1 ⍀ j2 ⍀ Figure P10.21 –j2 ⍀ ± – Vo 10 30° V 10 0° V ± – - 10.22 Find Vo in the circuit in Fig. P10.22. j1 ⍀ 1⍀ j2 ⍀ ± – 6 0° V Figure P10.26 + j2 ⍀ ± 1⍀ 4 30° A 1⍀ 10.27 Find Vo in the network in Fig. P10.27. j2 ⍀ – j1 ⍀ ± – 100 0° V ± – 200 0° V Vo – j5 ⍀ 10 ⍀ j2 ⍀ 2⍀ Vo –j10 ⍀ - Figure P10.22 Figure P10.27 irwin10_491-540hr.qxd 28-07-2010 12:10 Page 531 531 PROBLEMS 10.32 Determine the input impedance Z in of the circuit in 10.28 Determine the impedance seen by the source in the network shown in Fig. P10.28. Fig. P10.32. –j1 ⍀ j2 ⍀ 1⍀ 32 0° V ± – 1⍀ j2 ⍀ j2 ⍀ j1 ⍀ 1⍀ j1 ⍀ j2 ⍀ Zin 1⍀ j1 ⍀ j2 ⍀ 2⍀ –j2 ⍀ –j2 ⍀ Figure P10.28 Figure P10.32 10.29 Determine the impedance seen by the source in the network in Fig. P10.29. 10.33 Determine the input impedance Z in in the network in 6⍀ Fig. P10.33. j1 ⍀ j1⍀ 2⍀ 3⍀ j4 ⍀ 1⍀ Zin 12 0° V – + j4⍀ j2⍀ –j 1 ⍀ j4 ⍀ –j2 ⍀ 3⍀ –j1 ⍀ j2⍀ Figure P10.33 Figure P10.29 10.30 Determine the input impedance of the network shown in Fig. P10.30. j1⍀ –j 1 ⍀ whether a value of Xc can be found such that the output voltage is equal to twice the input voltage. –j 1 ⍀ 2⍀ 10.34 Analyze the network in Fig. P10.34 and determine j2 ⍀ 1⍀ Zin j3⍀ j2⍀ 1⍀ + j1 ⍀ V ± – 6⍀ I1 I2 j2 ⍀ –jXC ⍀ Vo - j4⍀ 3⍀ Figure P10.34 –j 2 ⍀ 10.35 Given the network shown in Fig. P10.35, determine the Figure P10.30 10.31 Determine the input impedance seen by the source in the circuit in Fig. P10.31. –j 1 ⍀ 2⍀ jM=j6 ⍀ –j 1 ⍀ 12 ⍀ j4⍀ j2⍀ 2 0° V value of the capacitor C that will cause the impedance seen by the 24 / 0° V voltage source to be purely resistive, f ⫽ 60 Hz. + – j3⍀ j6⍀ 18 ⍀ 6⍀ 10 ⍀ 24 0° V –j 4 ⍀ Figure P10.31 4⍀ 1 –––– jC j3⍀ ± – j1 ⍀ j50 ⍀ j6 ⍀ Figure P10.35 irwin10_491-540hr.qxd 532 28-07-2010 CHAPTER 10 12:10 Page 532 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10.36 Two coils in a network are positioned such that there is 100% coupling between them. If the inductance of one coil is 10 mH and the mutual inductance is 6 mH, compute the inductance of the other coil. 10.42 Find all currents and voltages in the network in Fig. P10.42. 1:2 I1 10.37 The currents in the magnetically coupled inductors shown in Fig. P10.37 are known to be i1(t) = 8 cos (377t - 20º) mA and i2(t) = 4 cos (377t - 50º) mA. The inductor values are L1 = 2H, L2 = 1H, and k = 0.6. Determine v1(t) and v2(t). 12 0° V I2 1⍀ ± – + + V1 V2 - - i1(t) j1 ⍀ Ideal + + L1 v1(t) 3⍀ L2 - Figure P10.42 v2(t) 10.43 Determine Vo in the circuit in Fig. P10.43. - i2(t) 1⍀ Figure P10.37 1⍀ 1:2 + – j1 ⍀ 10.38 Determine the energy stored in the coupled inductors 10 30° V in Problem 10.37 at t = 1 ms. ± – j2 ⍀ Vo - 10.39 The currents in the network in Fig. P10.39 are known to be i1(t) = 10 cos(377t - 30º) mA and i2(t) = 20 cos(377t - 45º) mA. The inductances are L1 = 2H, L2 = 2H, and k = 0.8. Determine v1(t) and v2(t). Figure P10.43 10.44 Determine I 1, I 2, V1, and V2 in the network in Fig. P10.44. M i1(t) i2(t) + v1(t) Ideal 1⍀ + L1 L2 v2(t) 12 30° V - - ± – I1 I2 2:1 + + V1 V2 - - Figure P10.39 1⍀ Ideal 10.40 Determine the energy stored in the coupled inductors Figure P10.44 in the circuit in Fig. P10.39 at t = 1 ms. 10.45 Determine I 1, I 2, V1, and V2 in the network in Fig. P10.45. 10.41 Determine Vo in the circuit in Fig. P10.41. 1⍀ 1:2 2⍀ 2⍀ + –j1 ⍀ 24 45° V ± – j3 ⍀ Vo 10 30° V ± – 1:2 + V1 V2 Ideal Figure P10.45 I2 + - Ideal Figure P10.41 –j1 ⍀ I1 4⍀ j4 ⍀ irwin10_491-540hr.qxd 28-07-2010 12:11 Page 533 PROBLEMS 10.46 Find Vo in the circuit in Fig. P10.46. 2:1 2⍀ + –j 2 ⍀ –j 2 ⍀ 24 0° V j2⍀ 2⍀ + – Vo 2⍀ 2⍀ – Ideal Figure P10.46 10.47 Find I o in the network in Fig. P10.47. 2:1 6⍀ –j 8 ⍀ 12 ⍀ 36 0° V 18 ⍀ 6⍀ + – 6⍀ –j 3 ⍀ –j 4 ⍀ Io Ideal Figure P10.47 10.48 Find Vo in the network in Fig. P10.48. –j 2 ⍀ 2:1 j4⍀ 2⍀ 36 0° V + – –j 4 ⍀ 6 0° A + 4⍀ Vo – Ideal Figure P10.48 10.49 Find I o in the circuit in Fig. P10.49. 6 0° V –j 2 ⍀ 1⍀ –+ 6 0° A 3⍀ 2⍀ 2:1 Io j2⍀ Ideal Figure P10.49 533 irwin10_491-540hr.qxd 534 28-07-2010 CHAPTER 10 12:11 Page 534 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10.50 Determine I 1, I 2, V1, and V2 in the network in Fig. P10.50. I1 1⍀ + + V1 V2 - - ± – 1 0° V 1⍀ 1:2 I2 2⍀ 1 0° A Ideal Figure P10.50 10.51 Find I o in the circuit in Fig. P10.51. 4 0° V 1⍀ 12 0° V j1⍀ –+ + – 1:2 Io 1⍀ 1⍀ –j 1 ⍀ –j 1 ⍀ 1⍀ 6 0° V 1⍀ Ideal Figure P10.51 10.52 Determine I 1, I 2, V1, and V2 in the network in Fig. P10.52. I1 1⍀ 24 30° V ± – I2 1:2 + 1⍀ –j1 ⍀ 1⍀ + j1 ⍀ V1 - V2 –j1 ⍀ Ideal Figure P10.52 10.53 Find Vo in the network in Fig. P10.53. 10.54 Determine I 1, I 2, V1, and V2 in the network in Fig. P10.54. Ideal 1:1 1⍀ –j1 ⍀ + V1 24 0° V ± – - I1 + + –j2 ⍀ V2 - 2⍀ ± – + V1 V2 - Ideal 2⍀ - Figure P10.53 3⍀ + 12 0° V Vo Figure P10.54 I2 1:4 4⍀ –j10 ⍀ 24 0° V ± – irwin10_491-540hr.qxd 28-07-2010 12:11 Page 535 PROBLEMS 535 10.55 Find I in the network in Fig. P10.55. 2:1 1⍀ 1⍀ 4⍀ – j1 ⍀ 2 –45° A j4 ⍀ – ± 1 60° V I Ideal Figure P10.55 10.56 Find the current I in the network in Fig. P10.56. 12 ⍀ 1:2 4⍀ ± – –j2 ⍀ 12 ⍀ 12 ⍀ –j4 ⍀ 120 0° V –j4 ⍀ I Ideal Figure P10.56 10.57 Find Vo in the network in Fig. P10.57. 10.58 Find Vo in the circuit in Fig. P10.58. –j1.2 ⍀ 2:1 4⍀ 32 0° V –j8 ⍀ ± – 1:2 + –j16 ⍀ 6⍀ Vo 1.6 ⍀ 10 ⍀ –j8 ⍀ Ideal .2. 1 2⍀ Ideal 32 0° V ± – 2⍀ j12 ⍀ – j2 ⍀ Figure P10.57 Figure P10.58 10.59 Find Vo in the circuit in Fig. P10.59. 1:2 –j2 ⍀ + –j1 ⍀ 1⍀ 1⍀ Vo Ideal - 2⍀ 24 0° V ± – Figure P10.59 2⍀ 2⍀ 2⍀ j2 ⍀ Vo - Ideal 6⍀ + 6⍀ irwin10_491-540hr.qxd 536 28-07-2010 CHAPTER 10 12:11 Page 536 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10.60 Find the voltage Vo in the network in Fig. P10.60. 2:1 1:4 j2 ⍀ – j2 ⍀ + 32 ⍀ 32 0° V ± – I2 Vo - I3 –j16 ⍀ I1 4⍀ Ideal Ideal Figure P10.60 10.61 Determine the input impedance seen by the source in the circuit in Fig. P10.61. 1:2 2⍀ VS ± – ± ± V1 V2 – – 4⍀ j2 ⍀ Ideal Figure P10.61 10.62 Determine the input impedance seen by the source in the circuit in Fig. P10.62. 1⍀ VS ± – I1 –j1 ⍀ 1⍀ 4:1 + + V1 V2 - - –j2 ⍀ 2⍀ Ideal Figure P10.62 10.63 Determine the input impedance seen by the source in the network shown in Fig. P10.63. –j1 ⍀ 2:1 j1 ⍀ 1:4 48 ⍀ 12 0° V ± – –j32 ⍀ 1⍀ Ideal Figure P10.63 Ideal irwin10_491-540hr.qxd 28-07-2010 12:11 Page 537 PROBLEMS 537 10.64 Determine the input impedance seen by the source in the network shown in Fig. P10.64. 4:1 2:1 20 ⍀ 4⍀ –j4 ⍀ 4⍀ VS ± – j2 ⍀ Ideal Ideal Figure P10.64 10.65 The output stage of an amplifier in an old radio is to be matched to the impedance of a speaker, as shown in Fig. P10.65. If the impedance of the speaker is 8 ⍀ and the amplifier requires a load impedance of 3.2 k⍀, determine the turns ratio of the ideal transformer. 10.67 Determine VS in the circuit in Fig. 10.67. 1:2 j1 ⍀ 1⍀ 1⍀ VS ± – + j1 ⍀ n:1 Vo=4 30° V Ideal Amplifier Figure P10.67 Ideal 10.68 Determine I S in the circuit in Fig. P10.68. Figure P10.65 1:2 10.66 Given that Vo = 48 / 30° V in the circuit shown in I1 6⍀ VS + + V1 V2 - 24 ⍀ Vo Ideal Figure P10.68 - Ideal Figure P10.66 10.69 In the network in Fig. P10.69, if I 1 = 4 / 0° A, find VS. 2:1 4⍀ 1:2 2⍀ –j1 ⍀ VS ± – – j4 ⍀ 1⍀ 8⍀ I1 Ideal Figure P10.69 Ideal I2=2 j1 ⍀ I2 1:2 + ± – 2⍀ IS –j6 ⍀ 1⍀ –j2 ⍀ Fig. P10.66, determine VS. 30° A irwin10_491-540hr.qxd 538 28-07-2010 CHAPTER 10 12:11 Page 538 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10.70 In the circuit in Fig. P10.70, if I x = 4 / 30° A, find Vo. + Vo – j8 ⍀ 1:2 2⍀ 1:2 2⍀ Ix 4⍀ –j1 ⍀ VS ± – 8⍀ 1⍀ Ideal Ideal Figure P10.70 10.71 For maximum power transfer, we desire to match the impedance of the inverting amplifier stage in Fig. P10.71 to the 50-⍀ equivalent resistance of the ac input source. However, standard op-amps perform best when the resistances around them are at least a few hundred ohms. The gain of the op-amp circuit should be -10. Design the complete circuit by selecting resistors no smaller than 1 k⍀ and specifying the turns ratio of the ideal transformer to satisfy both the gain and impedance matching requirements. R2 50 ⍀ 1:n + R1 – ± vin(t) + vo(t) - Ideal gain=–10 Figure P10.71 10.72 Digital clocks often divide a 60-Hz frequency signal to obtain a 1-second, 1-minute, or 1-hour signal. A convenient source of this 60-Hz signal is the power line. However, 120 volts is too high to be used by the low power electronics. Instead, a 3-V, 60-Hz signal is needed. If a resistive voltage divider is used to drop the voltage from 120 to 3 V, the heat generated will be unacceptable. In addition, it is costly to use a transformer in this application. Digital clocks are consumer items and must be very inexpensive to be a competitive product. The problem then is to design a circuit that will produce between 2.5 V and 3 V at 60 Hz from the 120-V ac power line without dissipating any heat or the use of a transformer. The design will interface with a circuit that has an input resistance of 1200 ohms. irwin10_491-540hr.qxd 28-07-2010 12:11 Page 539 TYPICAL PROBLEMS FOUND ON THE FE EXAM TYPICAL PROBLEMS FOUND ON THE FE EXAM • 10PFE-1 In the network in Fig. 10PFE-1, find the impedance seen by the source. a. 4.88⬔19.75°⍀ b. 2.56⬔31.26°⍀ c. 5.37⬔ -26.57°⍀ d. 8.23⬔ -10.61°⍀ M 4⍀ 24 cos (2t+0°) V ± – 100 mF 1H 5⍀ 4H k=0.5 Figure 10PFE-1 N2 N1 to achieve impedance matching for maximum power transfer. Using this value of n, calculate the power absorbed by the 3-⍀ resistor. 10PFE-2 In the circuit in Fig. 10PFE-2, select the value of the transformer’s turns ratio n = a. 100.75 W b. 37.5 W c. 55.6 W d. 75 W N1 : N2 48 ⍀ 120 0° V j32 ⍀ –j2 ⍀ ± – 3⍀ Ideal Figure 10PFE-2 10PFE-3 In the circuit in Fig. 10PFE-3, select the turns ratio of the ideal transformer that will match the output of the transistor amplifier to the speaker represented by the 16-⍀ load. a. 18 b. 30 c. 10 d. 25 a:1 1 k⍀ VS ± – 5 k⍀ + Vx 0.04 Vx 16 ⍀ (speaker) 10 k⍀ Ideal Figure 10PFE-3 539 irwin10_491-540hr.qxd 540 28-07-2010 CHAPTER 10 12:11 Page 540 M A G N E T I C A L LY C O U P L E D N E T W O R K S 10PFE-4 What is the current I 2 in the circuit shown in Fig. 10PFE-4? a. 11.77⬔35.25° A b. 5.85⬔20.62° A c. 23.54⬔11.31° A d. 15.36⬔8.48° A I1 120 0° V 6⍀ ± – j2 ⍀ I2 2:1 ± ± V1 V2 – – 1⍀ –j1 ⍀ Ideal Figure 10PFE-4 10PFE-5 What is the current I 2 in the circuit shown in Fig. 10PFE-5? a. 16.97⬔-45° A b. 10.54⬔30° A c. 12.02⬔-15° A d. 8.25⬔45° A I1 120 0° V ± – ± ± V1 V2 – – Ideal Figure 10PFE-5 I2 1:2 10 ⍀ j10 ⍀ irwin11_541-576hr.qxd 6-08-2010 15:11 Page 541 11 C H A PT E R POLYPHASE CIRCUITS THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Know the characteristics of a balanced, three-phase circuit ■ Know the basic wye and delta three-phase connections ■ Know how to calculate voltage and currents in balanced, three-phase circuits ■ Know how to calculate complex power in bal- anced, three-phase circuits Stephen Strathdee/iStockphoto W Wind Farms Turning a former waste site into a reliable source of renewable energy seems like a great idea—a real “win-win” for energy and the environment. Denmark has done just that impact of wind farms is remarkable—avoiding tons of chemical pollution that would occur from conventional energy sources. Renewable power generated by wind farms—despite the with its Middelgrunden Wind Farm on a natural reef east of the variability of wind itself—is added directly to the electric grid or Copenhagen harbor. Established in 2000 as the world’s stored in such media as electrochemical batteries or large- largest offshore wind farm, it consists of 20 turbines with a capacity capacitors. The final step is providing this power to total rated power of 40 Megawatts. For more than 200 years, the customer—a polyphase AC electrical system can deliver the reef had been a dumping site for harbor sludge and other more power at less voltage with smaller wires than the single- contaminated waste. Today, airline travelers flying into phase systems you considered earlier. We show in this chapter Copenhagen from the north can view the visual impact of this that a balanced three-phase system provides a constant instan- renewable energy source as a recognizable tribute to Danish taneous power to its load. Working on a per-phase basis allows ingenuity. you to easily calculate phase and line voltages and currents as Denmark has more than 6200 wind turbines producing an well as the power dissipated in the line and power delivered to estimated 89 million kilowatt-hours of electricity each year. One the load. High efficiency is essential both for wind farm power of Denmark’s largest industries, wind turbine manufacturing has generation and for power usage by three-phase systems—we a considerable export market worldwide. The environmental strive to minimize power losses in all cases. 541 irwin11_541-576hr.qxd 542 6-08-2010 CHAPTER 11 11.1 Three-Phase Circuits Figure 11.1 Hydroelectric generating facility. (Courtesy of Mark Nelms) 15:11 Page 542 P O LY P H A S E C I R C U I T S In this chapter we add a new dimension to our study of ac steady-state circuits. Up to this point we have dealt with what we refer to as single-phase circuits. Now we extend our analysis techniques to polyphase circuits or, more specifically, three-phase circuits (that is, circuits containing three voltage sources that are one-third of a cycle apart in time). We study three-phase circuits for a number of important reasons. It is more advantageous and economical to generate and transmit electric power in the polyphase mode than with single-phase systems. As a result, most electric power is transmitted in polyphase circuits. In the United States the power system frequency is 60 Hz, whereas in other parts of the world 50 Hz is common. The generation of electric power in the polyphase mode is accomplished with an electric generator, which converts mechanical energy to electrical energy. This mechanical energy can be produced at a dam or hydroelectric facility as shown in Fig. 11.1. As illustrated in Fig. 11.2, water stored in a reservoir falls through a turbine to the river below. The turbine drives the electric generator to produce three-phase voltages. In the fossil-fuel generating facility in Fig. 11.3, the turbine is driven by steam. In the diagram of Fig. 11.4, fuel and air are combusted in the boiler turning water into steam to drive the turbine. Cooling water is circulated through the condenser to change the steam exhaust from the turbine back to water to complete the cycle. A nuclear generating facility, shown in Fig. 11.5, also utilizes steam to drive the turbine. The heat from fission in the reactor core produces the steam. Note that all three types of generating facilities are located close to a body of water such as a river and are not often close to the loads that consume the electrical energy. Power transmission lines, such as those shown in Fig. 11.6, are constructed to transport electrical energy from the generating facilities to the loads. The transmission of electrical energy is most efficiently accomplished at very high voltages. Because this voltage can be extremely high in comparison to the level at which it is normally used (e.g., in the household), there is a need to raise and lower the voltage. This can be easily accomplished in ac systems using transformers, which we studied in Chapter 10. An example of a three-phase power transformer is shown in Fig. 11.7. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 543 SECTION 11.1 THREE-PHASE CIRCUITS 543 Reservoir Generator Turbine Figure 11.2 Diagram of a hydroelectric generating facility. (Diagram courtesy of Southern Company) Figure 11.3 A fossil-fuel generating facility. (Courtesy of Mark Nelms) irwin11_541-576hr.qxd 544 6-08-2010 CHAPTER 11 15:11 Page 544 P O LY P H A S E C I R C U I T S Figure 11.4 Conceptual diagram for a fossil-fuel generating facility. (Diagram courtesy of Southern Company) Generator Steam Turbine Electricity Steam exhaust Fuel Boiler Air Condenser Water Figure 11.5 A nuclear generating facility. (Stockbyte/SUPERSTOCK) Figure 11.6 Power transmission lines. (Courtesy of Mark Nelms) Cooling water irwin11_541-576hr.qxd 6-08-2010 15:11 Page 545 SECTION 11.1 THREE-PHASE CIRCUITS 545 Figure 11.7 A three-phase power transformer. (Courtesy of Jeremy Nelms, Talquin Electric Cooperative, Inc.) As the name implies, three-phase circuits are those in which the forcing function is a three-phase system of voltages. If the three sinusoidal voltages have the same magnitude and frequency and each voltage is 120° out of phase with the other two, the voltages are said to be balanced. If the loads are such that the currents produced by the voltages are also balanced, the entire circuit is referred to as a balanced three-phase circuit. A balanced set of three-phase voltages can be represented in the frequency domain as shown in Fig. 11.8a, where we have assumed that their magnitudes are 120 V rms. From the figure we note that Van = 120 / 0° V rms Vbn = 120 / -120° V rms 11.1 Vcn = 120 / -240° V rms = 120 / 120° V rms a ± – Van=120 0° V rms Figure 11.8 Balanced three-phase voltages. b ± – Vbn=120 –120° V rms c ± – Vcn=120 van vbn –240° V rms vcn t n (a) (b) irwin11_541-576hr.qxd 546 6-08-2010 CHAPTER 11 15:11 Page 546 P O LY P H A S E C I R C U I T S Our double-subscript notation is exactly the same as that employed in the earlier chapters; that is, Van means the voltage at point a with respect to point n. We will also employ the double-subscript notation for currents; that is, Ian is used to represent the current from a to n. However, we must be very careful in this case to describe the precise path, since in a circuit there will be more than one path between the two points. For example, in the case of a single loop the two possible currents in the two paths will be 180° out of phase with each other. The preceding phasor voltages can be expressed in the time domain as van(t) = 120 12 cos t V vbn(t) = 120 12 cos (t - 120°) V 11.2 vcn(t) = 120 12 cos (t - 240°) V These time functions are shown in Fig. 11.8b. Finally, let us examine the instantaneous power generated by a three-phase system. Assume that the voltages in Fig. 11.8 are van(t) = Vm cos t V vbn(t) = Vm cos (t - 120°) V 11.3 vcn(t) = Vm cos (t - 240°) V If the load is balanced, the currents produced by the sources are ia(t) = Im cos (t - ) A ib(t) = Im cos (t - - 120°) A 11.4 ic(t) = Im cos (t - - 240°) A The instantaneous power produced by the system is p(t) = pa(t) + pb(t) + pc(t) = Vm Im[cos t cos (t - ) + cos (t - 120°) cos (t - - 120°) 11.5 + cos (t - 240°) cos (t - - 240°)] Using the trigonometric identity, cos cos = 1 C cos ( - ) + cos ( + )D 2 11.6 Eq. (11.5) becomes p(t) = Vm Im Ccos + cos (2t - ) + cos 2 + cos (2t - - 240°) + cos + cos (2t - - 480°) D 11.7 which can be written as p(t) = Vm Im C3 cos + cos (2t - ) 2 + cos (2t - - 120°) + cos (2t - + 120°) D 11.8 There exists a trigonometric identity that allows us to simplify the preceding expression. The identity, which we will prove later using phasors, is cos + cos ( - 120°) + cos ( + 120°) = 0 11.9 If we employ this identity, the expression for the power becomes p(t) = 3 Vm Im cos W 2 11.10 irwin11_541-576hr.qxd 6-08-2010 15:11 Page 547 SECTION 11.2 547 THREE-PHASE CONNECTIONS Note that this equation indicates that the instantaneous power is always constant in time rather than pulsating, as in the single-phase case. Therefore, power delivery from a three-phase voltage source is very smooth, which is another important reason power is generated in three-phase form. 11.2 By far the most important polyphase voltage source is the balanced three-phase source. This source, as illustrated by Fig. 11.9, has the following properties. The phase voltages—that is, the voltage from each line a, b, and c to the neutral n—are given by Three-Phase Connections Van = Vp / 0° Vbn = Vp / -120° 11.11 Vcn = Vp / +120° Figure 11.9 Phase Balanced three-phase power source a Phase Van Balanced three-phase voltage source. a b Vbn b Phase Vcn c c n The phasor diagram for these voltages is shown in Fig. 11.10. The phase sequence of this set is said to be abc (called positive phase sequence), meaning that Vbn lags Van by 120°. We will standardize our notation so that we always label the voltages Van , Vbn , and Vcn and observe them in the order abc. Furthermore, we will normally assume with no loss of generality that / Van = 0°. An important property of the balanced voltage set is that Van + Vbn + Vcn = 0 11.12 This property can easily be seen by resolving the voltage phasors into components along the real and imaginary axes. It can also be demonstrated via Eq. (11.9). From the standpoint of the user who connects a load to the balanced three-phase voltage source, it is not important how the voltages are generated. It is important to note, however, that if the load currents generated by connecting a load to the power source shown in Fig. 11.9 are also balanced, there are two possible equivalent configurations for the load. The equivalent load can be considered as being connected in either a wye (Y) or a delta () configuration. The balanced wye configuration is shown in Fig. 11.11a and equivalently in Fig. 11.11b. The delta configuration is shown in Fig. 11.12a and equivalently in Fig. 11.12b. Note that in the case of the delta connection, there is no neutral line. The actual function of the neutral line in the wye connection will be examined, and it will be shown that in a balanced system the neutral line carries no current and, for purposes of analysis, may be omitted. The wye and delta connections each have their advantages. In the wye case, we have access to two voltages, the line-to-line and line-to-neutral, and it provides a convenient place to connect to ground for system protection. That is, it limits the magnitude of surge voltages. In the delta case, this configuration stays in balance better when serving unbalanced loads, and it is capable of trapping the third harmonic. Vcn 120° 120° 120° Van Vbn Figure 11.10 Phasor diagram for a balanced three-phase voltage source. irwin11_541-576hr.qxd 548 6-08-2010 CHAPTER 11 15:11 Page 548 P O LY P H A S E C I R C U I T S Figure 11.11 a Wye ( Y )-connected loads. b a ZY ZY b ZY c ZY ZY c n n Figure 11.12 Delta ()-connected loads. ZY Load Load (a) (b) a a Z⌬ Z⌬ Z⌬ Z⌬ b b Z⌬ Z⌬ c 11.3 + Vbn + Vcn + a Load Load (a) (b) Since the source and the load can each be connected in either Y or , three-phase balanced circuits can be connected Y–Y, Y–, –Y, or –. Our approach to the analysis of all of these circuits will be “Think Y”; therefore, we will analyze the Y–Y connection first. Source/Load Connections Van c Ia b Ib c Ic BALANCED WYE–WYE CONNECTION Suppose now that both the source and load are connected in a wye, as shown in Fig. 11.13. The phase voltages with positive phase sequence are Van = Vp / 0° Vbn = Vp / -120° ZY ZY ZY In Vcn = Vp / +120° where Vp , the phase voltage, is the magnitude of the phasor voltage from the neutral to any line. The line-to-line voltages or, simply, line voltages can be calculated using KVL; for example, Vab = Van - Vbn = Vp / 0° - Vp / -120° = Vp - Vp c - Figure 11.13 Balanced three-phase wye–wye connection. 11.13 = Vp c 1 13 - j d 2 2 3 13 + j d 2 2 = 13 Vp / 30° irwin11_541-576hr.qxd 6-08-2010 15:11 Page 549 SECTION 11.3 Vca Vcn Vcn Vab SOURCE/LOAD CONNECTIONS Figure 11.14 Vab Phasor representation of phase and line voltages in a balanced wye–wye system. 30° 30° Van Vbn Van Vbc Vbn (a) (b) The phasor addition is shown in Fig. 11.14a. In a similar manner, we obtain the set of lineto-line voltages as 11.14 Vca = 13 Vp / -210° All the line voltages together with the phase voltages are shown in Fig. 11.14b. We will denote the magnitude of the line voltages as VL , and therefore, for a balanced system, VL = 13 Vp 11.15 Hence, in a wye-connected system, the line voltage is equal to 13 times the phase voltage. As shown in Fig. 11.13, the line current for the a phase is Ia = Vp / 0° Van = ZY ZY 11.16 where Ib and Ic have the same magnitude but lag Ia by 120° and 240°, respectively. The neutral current In is then I n = AI a + I b + I c B = 0 11.17 Since there is no current in the neutral, this conductor could contain any impedance or it could be an open or a short circuit, without changing the results found previously. As illustrated by the wye–wye connection in Fig. 11.13, the current in the line connecting the source to the load is the same as the phase current flowing through the impedance ZY . Therefore, in a wye–wye connection, IL = IY [hint] Conversion rules: Vab = 13 Vp / 30° Vbc = 13 Vp / -90° 549 11.18 where IL is the magnitude of the line current and IY is the magnitude of the current in a wyeconnected load. Although we have a three-phase system composed of three sources and three loads, we can analyze a single phase and use the phase sequence to obtain the voltages and currents in the other phases. This is, of course, a direct result of the balanced condition. We may even have impedances present in the lines; however, as long as the system remains balanced, we need analyze only one phase. If the line impedances in lines a, b, and c are equal, the system will be balanced. Recall that the balance of the system is unaffected by whatever appears in the neutral line, and since the neutral line impedance is arbitrary, we assume that it is zero (i.e., a short circuit). / Vab = / Van + 30° Vab = 23 Van irwin11_541-576hr.qxd 550 6-08-2010 CHAPTER 11 15:11 Page 550 P O LY P H A S E C I R C U I T S EXAMPLE 11.1 SOLUTION An abc-sequence three-phase voltage source connected in a balanced wye has a line voltage of Vab = 208 / -30° V rms. We wish to determine the phase voltages. The magnitude of the phase voltage is given by the expression 208 13 = 120 V rms Vp = [hint] The phase of Van = / Van = / Vab - 30° The phase relationships between the line and phase voltages are shown in Fig. 11.14. From this figure we note that Van = 120 / -60° V rms Vbn = 120 / -180° V rms Vcn = 120 / +60° V rms The magnitudes of these voltages are quite common, and one often hears that the electric service in a building, for example, is three-phase 208/120 V rms. EXAMPLE 11.2 SOLUTION A three-phase wye-connected load is supplied by an abc-sequence balanced three-phase wye-connected source with a phase voltage of 120 V rms. If the line impedance and load impedance per phase are 1+j1 and 20+j10 , respectively, we wish to determine the value of the line currents and the load voltages. The phase voltages are Van = 120 / 0° V rms Vbn = 120 / -120° V rms Vcn = 120 / +120° V rms [hint] The per-phase circuit diagram is shown in Fig. 11.15. The line current for the a phase is / IbB = / IaA - 120° / IcC = / IaA + 120° I aA = 120 / 0° 21 + j11 = 5.06 / -27.65° A rms The load voltage for the a phase, which we call VAN, is VAN = A5.06 / -27.65°B(20 + j10) = 113.15 / -1.08° V rms Figure 11.15 Per-phase circuit diagram for the problem in Example 11.2. a IaA 1⍀ j1 ⍀ A 20 ⍀ ± Van – VAN j10 ⍀ n N irwin11_541-576hr.qxd 6-08-2010 15:11 Page 551 SECTION 11.3 SOURCE/LOAD CONNECTIONS 551 The corresponding line currents and load voltages for the b and c phases are I bB = 5.06 / -147.65° A rms VBN = 113.15 / -121.08° V rms I cC = 5.06 / -267.65° A rms VCN = 113.15 / -241.08° V rms To reemphasize and clarify our terminology, phase voltage, Vp , is the magnitude of the phasor voltage from the neutral to any line, while line voltage, VL , is the magnitude of the phasor voltage between any two lines. Thus, the values of VL and Vp will depend on the point at which they are calculated in the system. Learning Assessments E11.1 The voltage for the a phase of an abc-phase-sequence balanced wye-connected source is Van = 120 / 90° V rms. Determine the line voltages for this source. ANSWER: Vab = 208 / 120° V rms; Vbc = 208 / 0° V rms; Vca = 208 / -120° V rms. E11.2 An abc-phase-sequence three-phase voltage source connected in a balanced wye has a ANSWER: line voltage of Vab = 208 / 0° V rms. Determine the phase voltages of the source. Van = 120 / -30° V rms; Vbn = 120 / -150° V rms; Vcn = 120 / -270° V rms. E11.3 A three-phase wye-connected load is supplied by an abc-sequence balanced three-phase ANSWER: wye-connected source through a transmission line with an impedance of 1 + j1 per phase. The load impedance is 8 + j3 per phase. If the load voltage for the a phase is 104.02 / 26.6° V rms Ai.e., Vp = 104.02 V rms at the load endB, determine the phase voltages of the source. Van = 120 / 30° V rms; E11.4 A positive-sequence balanced three-phase wye-connected source with a phase voltage of 277 V rms supplies power to a balanced wye-connected load. The per-phase load impedance is 60 - j40 . Determine the line currents in the circuit if the phase angle of Van = 0°. ANSWER: IaA = 3.84 / 33.69° A rms; IbB = 3.84 / -86.31° A rms; IcC = 3.84 / 153.69° A rms. E11.5 An abc-sequence set of voltages feeds a balanced three-phase wye–wye system. ANSWER: Vab = 214.8 / 41.6° V rms; Vbc = 214.8 / -78.4° V rms; Vca = 214.8 / 161.6° V rms. The line and load impedances are 0.5 + j0.75 and 20 - j24 respectively. If the load voltage of the a-phase is VAN = 125 / 10° V rms, find the line voltages of the input. E11.6 In a balanced three-phase wye–wye system, the total power in the lines is 650 W. VAN = 117 / 15° V rms and the power factor of the load is 0.88 leading. If the line impedance is 1 + j2 , determine the load impedance. Vbn = 120 / -90° V rms; Vcn = 120 / -210° V rms. ANSWER: ZL = 7 - j3.78 . irwin11_541-576hr.qxd 552 6-08-2010 CHAPTER 11 15:11 Page 552 P O LY P H A S E C I R C U I T S The previous analysis indicates that we can simply treat a three-phase balanced circuit on a per-phase basis and use the phase relationship to determine all voltages and currents. Let us now examine the situations in which either the source or the load is connected in . DELTA-CONNECTED SOURCE Consider the delta-connected source shown in Fig. 11.16a. Note that the sources are connected line to line. We found earlier that the relationship between line-to-line and line-to-neutral voltages was given by Eq. (11.14) and illustrated in Fig. 11.14 for an abc-phase sequence of voltages. Therefore, if the delta sources are Vab = VL / 0° Vbc = VL / -120° 11.19 Vca = VL / +120° where VL is the magnitude of the phase voltage. The equivalent wye sources shown in Fig. 11.16b are Van = Vbn = Vcn = VL 13 VL 13 VL 13 / -30° = Vp / -30° / -150° = Vp / -150° 11.20 / -270° = Vp / +90° where Vp is the magnitude of the phase voltage of an equivalent wye-connected source. Therefore, if we encounter a network containing a delta-connected source, we can easily convert the source from delta to wye so that all the techniques we have discussed previously can be applied in an analysis. Ia a a Sources connected in delta and wye. Vca + + b Ic (a) Ib Vcn c + c ± Van – + V - ab Vbc Ia + Figure 11.16 Vbn b Ib Ic (b) Problem-Solving Strategy Three-Phase Balanced AC Power Circuits Step 1. Convert the source/load connection to a wye–wye connection if either the source, load, or both are connected in delta since the wye–wye connection can be easily used to obtain the unknown phasors. Step 2. Only the unknown phasors for the a-phase of the circuit need be determined since the three-phase system is balanced. Step 3. Finally, convert the now known phasors to the corresponding phasors in the original system. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 553 SECTION 11.3 SOURCE/LOAD CONNECTIONS Consider the network shown in Fig. 11.17a. We wish to determine the line currents and the magnitude of the line voltage at the load. The single-phase diagram for the network is shown in Fig. 11.17b. The line current IaA is I aA = 553 EXAMPLE SOLUTION 11.3 A208兾 13B / -30° 12.1 + j4.2 = 9.38 / -49.14° A rms and thus I bB = 9.38 / -169.14° V rms and I cC = 9.38 / 70.86° V rms. The voltage VAN is then VAN = A9.38 / -49.14°B(12 + j4) = 118.65 / -30.71° V rms Therefore, the magnitude of the line voltage at the load is VL = 13 (118.65) = 205.51 V rms The phase voltage at the source is Vp = 208兾 13 = 120 V rms, while the phase voltage at the load is Vp = 205.51兾13 = 118.65 V rms. Clearly, we must be careful with our notation and specify where the phase or line voltage is taken. IaA a a 208 –240° V rms + ± – b ± – 0.1 ⍀ j0.2 ⍀ A 12 ⍀ Delta–wye network and an equivalent single-phase (a-phase) diagram. 0.1 ⍀ j0.2 ⍀ 12 ⍀ j0.2 ⍀ B 12 ⍀ j4 ⍀ N ± – 208 –30° V rms 13 ––– VAN j4 ⍀ 208 –120° V rms 0.1 ⍀ 12 ⍀ j0.2 ⍀ c A j4 ⍀ 208 0° V rms 0.1 ⍀ Figure 11.17 j4 ⍀ n C N (a) (b) Learning Assessments E11.7 Consider the network shown in Fig. E11.7. Compute the magnitude of the line voltages at the load. a ± – ± – 208 –240° V rms 0.1 ⍀ ± – c 10 ⍀ j4 ⍀ j0.1 ⍀ B 10 ⍀ j4 ⍀ j0.1 ⍀ C 10 ⍀ j4 ⍀ 208 –120° V rms 0.1 ⍀ Figure E11.7 A 208 0° V rms 0.1 ⍀ b j0.1 ⍀ N ANSWER: VL = 205.2 V rms. irwin11_541-576hr.qxd 554 6-08-2010 CHAPTER 11 15:11 Page 554 P O LY P H A S E C I R C U I T S ANSWER: VL = 209.2 V rms. E11.8 Find the magnitude of the line voltage at the load in Fig. E11.8. a j0.1 0.05 A 9 + – 9 215 –10° V rms j6 215 110° V rms – + j 0.1 0.05 b B 9 + – j6 215 –130° V rms j6 j 0.1 0.05 c Figure E11.8 C Consider now the -connected load shown in Fig. 11.18. Note that in this connection the line-to-line voltage is the voltage across each load impedance. If the phase voltages of the source are DELTA-CONNECTED LOAD Van = Vp / 0° Vbn = Vp / -120° 11.21 Vcn = Vp / +120° then the line voltages are Vab = 13 Vp / 30° = VL / 30° = VAB Vbc = 13 Vp / -90° = VL / -90° = VBC Vca = 13 Vp / -210° = VL / -210° = VCA Figure 11.18 Van Balanced three-phase wye–delta system. a IaA A ±– IAB Vab Vbn ±– b IbB ZΔ Vcn Vbc IcC c ICA B ZΔ IBC ZΔ ±– n Vca C 11.22 irwin11_541-576hr.qxd 6-08-2010 15:11 Page 555 SECTION 11.3 SOURCE/LOAD CONNECTIONS 555 where VL is the magnitude of the line voltage at both the delta-connected load and at the source since there is no line impedance present in the network. From Fig. 11.18 we note that if Z ¢ = Z¢ / , the phase currents at the load are I AB = VAB Z¢ 11.23 where IBC and ICA have the same magnitude but lag IAB by 120° and 240°, respectively. KCL can now be employed in conjunction with the phase currents to determine the line currents. For example, I aA = I AB + I AC = I AB - I CA However, it is perhaps easier to simply convert the balanced -connected load to a balanced Y-connected load using the –Y transformation. This conversion is possible since the wye–delta and delta–wye transformations outlined in Chapter 2 are also valid for impedance in the frequency domain. In the balanced case, the transformation equations reduce to ZY = 1 Z¢ 3 and then the line current IaA is simply I aA = Van ZY Finally, using the same approach as that employed earlier to determine the relationship between the line voltages and phase voltages in a Y–Y connection, we can show that the relationship between the magnitudes of the phase currents in the -connected load and the line currents is IL = 13 I¢ 11.24 A balanced delta-connected load contains a 10- resistor in series with a 20-mH inductor in each phase. The voltage source is an abc-sequence three-phase 60-Hz, balanced wye with a voltage Van = 120 / 30° V rms. We wish to determine all currents and line currents. The impedance per phase in the delta load is Z=10+j7.54 . The line voltage Vab = 120 13 / 60° V rms. Since there is no line impedance, VAB = Vab = 120 13 / 60° V rms. Hence, 120 13 / 60° I AB = 10 + j7.54 = 16.60 / +22.98° A rms If Z = 10 + j7.54 , then ZY = 1 Z 3 ¢ = 3.33 + j2.51 Then the line current I aA = 120 / 30° Van = ZY 3.33 + j2.51 120 / 30° = 4.17 / 37.01° = 28.78 / -7.01° A rms EXAMPLE 11.4 SOLUTION irwin11_541-576hr.qxd 556 6-08-2010 CHAPTER 11 15:11 Page 556 P O LY P H A S E C I R C U I T S Therefore, the remaining phase and line currents are I BC = 16.60 / -97.02° A rms I bB = 28.78 / -127.01° A rms I CA = 16.60 / +142.98° A rms I cC = 28.78 / +112.99° A rms In summary, the relationship between the line voltage and phase voltage and the line current and phase current for both the Y and configurations are shown in Fig. 11.19. The currents and voltages are shown for one phase. The two remaining phases have the same magnitude but lag by 120° and 240°, respectively. IL=IL + Vp VL=13 Vp a or A IL + +30° IL=IL - a or A + + n VL=VL +30° b or B IL ––– - - - c or C 13 +30° c or C b or B (a) (b) Figure 11.19 Voltage and current relationships for Y and configurations. Careful observation of Table 11.1 indicates that the following rules apply when solving problems in balanced three-phase systems: ● ● ● ● The phase of the voltages and currents in a connection is 30° ahead of those in a Y connection. The magnitude of the line voltage or, equivalently, the -connection phase voltage, is 13 times that of the Y-connection phase voltage. The magnitude of the line current or, equivalently, the Y-connection phase current, is 13 times that of the -connection phase current. The load impedance in the Y connection is one-third of that in the -connection, and the phase is identical. TABLE 11.1 The voltage, current, and impedance relationships for Y and configurations Y Line voltage AVab or VAB B 13 Vp / + 30° VL / + 30° = VL / + 30° Line current IaA IL / IL / Phase voltage Vp / AVan or VAN B 13 Vp / + 30° Phase current IL / Load impedance ZY / - IL 13 / + 30° 3 ZY / - irwin11_541-576hr.qxd 6-08-2010 15:11 Page 557 SECTION 11.4 P O W E R R E L AT I O N S H I P S 557 Learning Assessments E11.9 An abc-sequence three-phase voltage source connected in a balanced wye supplies power to a balanced delta-connected load. The line current for the a phase is I aA = 12 / 40° A rms. Find the phase currents in the delta-connected load. ANSWER: I AB = 6.93 / 70° A rms; I BC = 6.93 / -50° A rms; I CA = 6.93 / -170° A rms. E11.10 Find the line currents and the power absorbed by the delta-connected load in Fig. E11.10. a 0.5 j1 A 30 + 480 0° V rms – 480 120° V rms – + 0.5 30 – j12 j1 B b 30 + 480 –120° V rms – 0.5 ANSWER: IaA = 35.76 / -34.74° A rms; IbB = 35.76 / -154.74° A rms; IcC = 35.76 / 85.26° A rms; 17.29 - j6.92 kVA. –j12 – j12 j1 C c 15 15 15 j8 j8 j8 Figure E11.10 Whether the load is connected in a wye or a delta, the real and reactive power per phase is Pp = Vp Ip cos Qp = Vp Ip sin 11.25 where is the angle between the phase voltage and the line current. For a Y-connected system, Ip = IL and Vp = VL兾13 , and for a -connected system, Ip = IL兾13 and Vp = VL . Therefore, Pp = Qp = VL IL 13 VL IL 13 cos 11.26 sin The total real and reactive power for all three phases is then PT = 3 PP = 13 VL IL cos QT = 3 QP = 13 VL IL sin and, therefore, the magnitude of the complex power (apparent power) is ST = 2P2T + Q2T = 13 VL IL and / ST = 11.4 11.27 Power Relationships irwin11_541-576hr.qxd 558 6-08-2010 CHAPTER 11 EXAMPLE 11.5 SOLUTION 15:11 Page 558 P O LY P H A S E C I R C U I T S A three-phase balanced wye–delta system has a line voltage of 208 V rms. The total real power absorbed by the load is 1200 W. If the power factor angle of the load is 20° lagging, we wish to determine the magnitude of the line current and the value of the load impedance per phase in the delta. The line current can be obtained from Eq. (11.26). Since the real power per phase is 400 W, 400 = 208IL cos 20° 13 IL = 3.54 A rms The magnitude of the current in each leg of the delta-connected load is IL I¢ = 13 = 2.05 A rms Therefore, the magnitude of the delta impedance in each phase of the load is @Z ¢ @ = = VL I¢ 208 2.05 = 101.46 Since the power factor angle is 20° lagging, the load impedance is Z ¢ = 101.46 / 20° = 95.34 + j34.70 EXAMPLE 11.6 SOLUTION For the circuit in Example 11.2 we wish to determine the real and reactive power per phase at the load and the total real power, reactive power, and complex power at the source. From the data in Example 11.2 the complex power per phase at the load is Sload = VI* = A113.15 / -1.08°BA5.06 / 27.65°B = 572.54 / 26.57° = 512.07 + j256.09 VA Therefore, the real and reactive power per phase at the load are 512.07 W and 256.09 var, respectively. The complex power per phase at the source is Ssource = VI* = A120 / 0°B A5.06 / 27.65°B = 607.2 / 27.65° = 537.86 + j281.78 VA Therefore, total real power, reactive power, and apparent power at the source are 1613.6 W, 845.2 var, and 1821.6 VA , respectively. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 559 SECTION 11.4 P O W E R R E L AT I O N S H I P S A balanced three-phase source serves three loads, as follows: Load 1: 24 kW at 0.6 lagging power factor Load 2: 10 kW at unity power factor Load 3: 12 kVA at 0.8 leading power factor If the line voltage at the loads is 208 V rms at 60 Hz, we wish to determine the line current and the combined power factor of the loads. From the data we find that 559 EXAMPLE 11.7 SOLUTION S1 = 24,000 + j32,000 S2 = 10,000 + j0 S3 = 12,000 / -36.9° = 9600 - j7200 Therefore, Sload = 43,600 + j24,800 = 50,160 / 29.63° VA [hint] The sum of three complex powers; Sload = S1 + S2 + S3 @Sload @ IL = 13 VL 50,160 = 208 13 IL = 139.23 A rms and the combined power factor is pfload = cos 29.63° = 0.869 lagging Given the three-phase system in Example 11.7, let us determine the line voltage and power factor at the source if the line impedance is Zline = 0.05 + j0.02 . The complex power absorbed by the line impedances is EXAMPLE SOLUTION 11.8 Sline = 3ARline I 2L + jXline I 2L B = 2908 + j1163 VA The complex power delivered by the source is then SS = Sload + Sline = 43,600 + j24,800 + 2908 + j1163 = 53,264 / 29.17° VA The line voltage at the source is then VLS = SS 13 IL = 220.87 V rms and the power factor at the source is pfS = cos 29.17° = 0.873 lagging [hint] Recall that the complex power for all three lines is Sline = 3I2LZline irwin11_541-576hr.qxd 560 6-08-2010 CHAPTER 11 15:11 Page 560 P O LY P H A S E C I R C U I T S EXAMPLE 11.9 SOLUTION Let’s consider the three-phase system shown in Fig. 11.20. Calculate the real power loss in the line resistance for VL = 500 kV rms and 50 kV rms. Sload For VL = 500 kV rms , IL = 23VL = 1000 23(500) = 1.155 kA rms , and the real power losses in the line are Pline = 3I 2LRline = 3(1.155)2(0.1) = 0.4 MW . For VL = 50 kV rms , IL = 1000 23(50) = 11.55 kA rms and Pline = 3I 2LRline = 3(11.55)2(0.1) = 40 MW The line losses at 50 kV rms are 100 times larger than those at 500 kV rms. This example illustrates that power transmission at higher voltages is more efficient because of the reduced losses. The transformer discussed in Chapter 10 allows voltage levels in ac systems to be changed easily. Electric generators at power plants generate line voltages up to 25 kV. Transformers are utilized to step up this voltage for transmission from the plants to the load centers. a Figure 11.20 Three-phase system for calculation of line losses for different load voltages. Balanced three-phase source b c 0.1 j0.2 0.1 j0.2 0.1 j0.2 + A - B VL C 1000 MVA 0.8 lagging LearningAssessments E11.11 A three-phase balanced wye–wye system has a line voltage of 208 V rms. The total real power absorbed by the load is 12 kW at 0.8 pf lagging. Determine the per-phase impedance of the load. E11.12 For the balanced wye–wye system described in Learning Assessment E11.3, determine the real and reactive power and the complex power at both the source and the load. ANSWER: Z = 2.88 / 36.87°. ANSWER: Sload = 1186.77 + j444.66 VA; Ssource = 1335.65 + j593.55 VA. E11.13 A 480-V rms line feeds two balanced three-phase loads. If the two loads are rated as follows, Load 1: 5 kVA at 0.8 pf lagging Load 2: 10 kVA at 0.9 pf lagging determine the magnitude of the line current from the 480-V rms source. ANSWER: IL = 17.97 A rms. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 561 SECTION 11.5 P OW E R FAC TO R C O R R E C T I O N E11.14 If the line voltage at the load is 480 V rms in Fig. E11.14, find the line voltage and power factor at the source. a Balanced three-phase source 0.1 561 ANSWER: VL = 501.7 V rms, pf = 0.9568 lagging. j0.2 A b 0.1 j0.2 c 0.1 j0.2 + 480 V rms – B C 40 kW 0.8 lagging 30 kVA 0.9 leading Figure E11.14 In Section 9.7 we illustrated a simple technique for raising the power factor of a load. The method involved judiciously selecting a capacitor and placing it in parallel with the load. In a balanced three-phase system, power factor correction is performed in exactly the same manner. It is important to note, however, that the Scap specified in Eq. (9.37) is provided by three capacitors, and in addition, Vrms in the equation is the voltage across each capacitor. The following example illustrates the technique. 11.5 Power Factor Correction [hint] Major precautions for three-phase power factor correction: Must distinguish PT and PP . Must use appropriate V rms for Y- and -connections. In the balanced three-phase system shown in Fig. 11.21, the line voltage is 34.5 kV rms at 60 Hz. We wish to find the values of the capacitors C such that the total load has a power factor of 0.94 leading. Following the development outlined in Section 9.7 for single-phase power factor correction, we obtain EXAMPLE 11.10 SOLUTION Sold = 24 / cos-1 0.78 MVA = 18.72 + j15.02 MVA and new = -cos-1 0.94 [hint] = -19.95° Therefore, Snew = 18.72 + j18.72 tan (-19.95°) = 18.72 - j6.80 MVA and Scap = Snew - Sold = -j21.82 MVA The reactive power to be supplied by C is derived from the expression jQcap = -jCV 2rms The phase voltage for the Y connection is 34.5k VY = 13 irwin11_541-576hr.qxd 562 6-08-2010 CHAPTER 11 15:11 Page 562 P O LY P H A S E C I R C U I T S However, -jC V2rms = -j21.82 MVA and since the line voltage is 34.5 kV rms, then (377) a 21.82 34.5k 2 M b C = 3 13 Hence, C = 48.6 F Figure 11.21 a Network used in Example 11.10. Balanced three-phase source Balanced load 24 MVA 0.78 power factor lagging b c C C C Neutral LearningAssessments E11.15 Find C in Example 11.10 such that the load has a power factor of 0.90 lagging. ANSWER: C = 13.26 F. E11.16 Find C in Fig. E11.16 such that the power factor of the source if 0.98 lagging. ANSWER: C = 14/ F. 13.8 kV rms Balanced three-phase source 60 Hz C C C Balanced three-phase load 6 MVA 0.75 lagging Figure E11.16 Finally, recall that our entire discussion in this chapter has focused on balanced systems. It is extremely important, however, to point out that in an unbalanced three-phase system the problem is much more complicated because of the mutual inductive coupling between phases in power apparatus. 11.6 Application Examples The first of the following three examples illustrates the manner in which power flow is measured when utilities are interconnected, answering the question of who is supplying power to whom. The last example demonstrates the actual method in which capacitors are specified by the manufacturer for power factor correction. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 563 SECTION 11.6 A P P L I C AT I O N E X A M P L E S 563 Two balanced three-phase systems, X and Y, are interconnected with lines having impedance Zline = 1 + j2 . The line voltages are Vab = 12 / 0° kV rms and VAB = 12 / 5° kV rms, as shown in Fig. 11.22a. We wish to determine which system is the source, which is the load, and the average power supplied by the source and absorbed by the load. APPLICATION EXAMPLE 11.11 When we draw the per phase circuit for the system as shown in Fig. 11.22b, the analysis will be essentially the same as that of Example 9.12. The network in Fig. 11.22b indicates that SOLUTION Van - VAN Z line 12,000 12,000 / -25° / -30° 13 13 = 15 / 63.43° I aA = = 270.30 / -180.93° A rms The average power absorbed by system Y is PY = 13 VAB IaA cos (Van - IaA) = 13 (12,000)(270.30) cos (-25° + 180.93°) = -5.130 MW Note that system Y is not the load, but rather the source and supplies 5.130 MW. System X absorbs the following average power: PX = 13 Vab IAa cos AVan - IaA B where I Aa = -I aA = 270.30 / -0.93° A rms Therefore, PX = 13 (12,000)(270.30) cos (-30° + 0.93°) = 4.910 MW and hence system X is the load. The difference in the power supplied by system Y and that absorbed by system X is, of course, the power absorbed by the resistance of the three lines. + b Vab - System X j2 ⍀ 1⍀ a c 1⍀ Zline Figure 11.22 A + j2 ⍀ VAB - B System 1⍀ j2 ⍀ n C Y N (a) a IaA j2 ⍀ 1⍀ + A + 12 Van=––– –30° kV rms 13 -n 12 VAN=––– –25° kV rms 13 - N (b) Circuits used in Example 11.11: (a) original three-phase system, (b) per-phase circuit. irwin11_541-576hr.qxd 564 6-08-2010 CHAPTER 11 15:11 Page 564 P O LY P H A S E C I R C U I T S The preceding example illustrates an interesting point. Note that the phase difference between the two ends of the power line determines the direction of the power flow. Since the numerous power companies throughout the United States are tied together to form the U.S. power grid, the phase difference across the interconnecting transmission lines reflects the manner in which power is transferred between power companies. Capacitors for power factor correction are usually specified by the manufacturer in vars rather than in farads. Of course, the supplier must also specify the voltage at which the capacitor is designed to operate, and a frequency of 60 Hz is assumed. The relationship between capacitance and the var rating is QR = V2 ZC where QR is the var rating, V is the voltage rating, and ZC is the capacitor’s impedance at 60 Hz. Thus, a 500-V, 600-var capacitor has a capacitance of C = QR V2 = 600 (377)(500)2 or C = 6.37 F and can be used in any application where the voltage across the capacitor does not exceed the rated value of 500 V. APPLICATION EXAMPLE 11.12 Let us examine, in a general sense, the incremental cost of power factor correction; specifically, how much capacitance is required to improve the power factor by a fixed amount, say 0.01? SOLUTION The answer to this question depends primarily on two factors: the apparent power and the original power factor before correction. This dependence can be illustrated by developing equations for the old and new power factors, and their corresponding power factor angles. We know that pfold = cosAold B tan Aold B = pfnew = cos Anew B Qold - QC tan Anew B = P Qold P 11.28 If the difference in the power factors is 0.01, then pfnew - pfold = 0.01 11.29 Solving for the ratio QC兾P, since reactive power and capacitance are proportional to one another will yield the reactive power per watt required to improve the power factor by 0.01. Using Eq. (11.28), we can write QC Qold = - tan AnewB = tan Aold B - tan AnewB = tan Cacos ApfoldBD - tan Cacos Apfold + 0.01BD P P 11.30 A plot of Eq. (11.30), shown in Fig. 11.23, has some rather interesting implications. First, the improvement required for a power factor change of 0.01 is at a minimum when the original power factor is about 0.81. Thus, an incremental improvement at that point is least expensive. Second, as the original power factor approaches unity, changes in power factor are more expensive to implement. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 565 Required reactive power per watt, QC /P SECTION 11.6 A P P L I C AT I O N E X A M P L E S 565 Figure 11.23 0.16 A plot of required reactive power per watt needed to improve the original power factor by 0.01. 0.12 0.08 0.04 0 0.4 0.5 0.6 0.7 0.8 0.9 1 Original power factor, pfold Table 11.2 lists the voltage and power ratings for three power factor correction capacitors. Let us determine which of them, if any, can be employed in Example 11.10. TABLE 11.2 APPLICATION EXAMPLE 11.13 Rated voltage and vars for power factor correction capacitors CAPACITOR RATED VOLTAGE (kV) RATED Q (Mvars) 1 2 3 10.0 50.0 20.0 4.0 25.0 7.5 From Fig. 11.21 we see that the voltage across the capacitors is the line-to-neutral voltage, which is Van = Vab 13 34,500 = 13 or Van = 19.9 kV Therefore, only those capacitors with rated voltages greater than or equal to 19.9 kV can be used in this application, which eliminates capacitor 1. Let us now determine the capacitance of capacitors 2 and 3. For capacitor 2, C2 = Q V 2 = 25 * 106 (377)(50,000)2 or C2 = 26.53 F which is much smaller than the required 48.6 F. The capacitance of capacitor 3 is C3 = Q V2 = 7.5 * 106 (377)(20,000)2 or C3 = 49.7 F which is within 2.5% of the required value. Obviously, capacitor 3 is the best choice. SOLUTION irwin11_541-576hr.qxd 566 6-08-2010 CHAPTER 11 15:11 Page 566 P O LY P H A S E C I R C U I T S 11.7 In the first example in this section, we examine the selection of both the conductor and the capacitor in a practical power factor correction problem. Design Examples DESIGN EXAMPLE 11.14 SOLUTION Two stores, as shown in Fig. 11.24, are located at a busy intersection. The stores are fed from a balanced three-phase 60-Hz source with a line voltage of 13.8 kV rms. The power line is constructed of a #4ACSR (aluminum cable steel reinforced) conductor that is rated at 170 A rms. A third store, shown in Fig. 11.24, wishes to locate at the intersection. Let us determine (1) if the #4ACSR conductor will permit the addition of this store, and (2) the value of the capacitors connected in wye that are required to change the overall power factor for all three stores to 0.92 lagging. 1. The complex power for each of the three loads is S1 = 700 / 36.9° = 560 + j420 kVA S2 = 1000 / 60° = 500 + j866 kVA S3 = 800 / 25.8° = 720 + j349 kVA Therefore, the total complex power is ST = S1 + S2 + S3 = 1780 + j1635 = 2417 / 42.57° kVA Since ST = 13 VL IL the line current is IL = (2417)A103 B 13 (13.8)A103 B = 101.1 A rms Since this value is well below the rated value of 170 A rms, the conductor is sized properly and we can safely add the third store. 2. The combined power factor for the three loads is found from the expression cos = pf = Figure 11.24 Circuit used in Example 11.14. Balanced three-phase source 13.8 kV 1780 = 0.7365 lagging 2417 a b c Store #1 700 kVA pf=0.8 lagging Store #2 1000 kVA pf=0.5 lagging Store #3 800 kVA pf=0.9 lagging irwin11_541-576hr.qxd 6-08-2010 15:11 Page 567 SECTION 11.7 DESIGN EXAMPLES 567 By adding capacitors we wish to change this power factor to 0.92 lagging. This new power factor corresponds to a new of 23.07°. Therefore, the new complex power is Snew = 1780 + j1780 tan (23.07°) = 1780 + j758.28 kVA As illustrated in Fig. 9.17, the difference between Snew and ST is that supplied by the purely reactive capacitor and, therefore, Scap = jQC = Snew - ST or jQC = j(758.28 - 1635) = -j876.72 kVA Thus, -jC V2rms = and 377 a -j876.72k 3 13.8 * 103 2 876.72 b C = * 103 3 13 Therefore, C = 12.2 F Hence, three capacitors of this value connected in wye at the load will yield a total power factor of 0.92 lagging. Control circuitry for high-voltage, three-phase equipment usually operates at much lower voltages. For example, a 10-kW power supply might operate at a line voltage of 480 V rms, while its control circuit is powered by internal dc power supplies at ;5 V. Lower voltages are not only safer to operate but also permit engineers to easily incorporate op-amps and digital electronics into the control system. It is a great convenience to test the control circuit without having to connect it directly to a 480-V rms three-phase source. Therefore, let us design a low-power, three-phase emulator that simulates a three-phase system at low voltage and low cost and provides a test bed for the control circuitry. The emulator should generate proper phasing but with a magnitude that is adjustable between 1 and 4 volts peak. DESIGN EXAMPLE 11.15 Our design, shown in Fig. 11.25, will consist of three parts: a magnitude adjustment, a phase angle generator, and a phase B generator. The ac input is a 60-Hz sine wave with a peak of about 5 V. This voltage can be generated from a standard 120 V rms wall outlet using a stepdown transformer with a turns ratio of SOLUTION n = 120 12 = 34 : 1 5 Figure 11.25 Magnitude adjustor 5 0° V ± – Phase Phase VBN angle + generator + generator + VAN VCN VBN - - - The potentiometer circuit shown in Fig. 11.26a can be used to provide magnitude adjustment. Resistors R1 and R2 provide the voltage limits of 1 and 4 V. We can use simple A block diagram for a threephase emulator. irwin11_541-576hr.qxd 568 6-08-2010 CHAPTER 11 15:11 Page 568 P O LY P H A S E C I R C U I T S Figure 11.26 Subcircuits within the threephase emulator: (a) the magnitude adjustor, (b) the R-C portion of the phase angle generator, (c) the complete phase angle generator, and (d) the generator for phase VBN. + R1 R3 5 sin (t) V ± – Rp V1 + + 1 V1 R2 –––– jC V2 - - - (a) (b) + R3 R5 V1 + 1 –––– jC - R4 ± – – ± + V2 V3 - (c) R6 R8 R7 – ± + VAN + - + VBN VCN - - (d) voltage division to determine the relationships between R1 , R2, and Rp . When the pot’s wiper arm is at the bottom of the pot in Fig. 11.26a, we have V1 = 1 = 5 c R2 d R1 + R2 + Rp 1 R1 + Rp = 4R2 11.31 1 R2 + Rp = 4R1 11.32 and when the wiper is at the top, V1 = 4 = 5 c R2 + Rp R1 + R2 + Rp d Solving Eqs. (11.31) and (11.32) yields the requirements that R1 = R2 = Rp兾3. To obtain values for these resistors we must simply choose one of them. We know that resistors are available in a wide variety of values in small increments. Potentiometers on the other hand are not. Typical potentiometer values are 10, 20, 50, 100, 200, 500, ...10k ... 100k, 200k, 500k, ... up to about 10 M. Since the potentiometer offers fewer options, we will choose its value to be 10 k, which yields R1 = R2 = 3.3 k—a standard resistor value. We will set V1 as the phase voltage VAN . irwin11_541-576hr.qxd 6-08-2010 15:11 Page 569 SECTION 11.7 DESIGN EXAMPLES 569 Next we consider the phase angle generator. Since capacitors are generally smaller physically than inductors, we will use the simple RC network in Fig. 11.26b to shift the phase of V1 . Assigning a phase angle of 0° to V1 , we know that the phase of V2 must be between 0 and -90 degrees. Unfortunately, in order to generate VCN we need a phase angle of +120°. If we create a phase angle of -60° and invert the resulting sine wave, we will produce an equivalent phase angle of +120°! The inversion can be performed by an inverting op-amp configuration. To produce a -60° phase angle at V2 requires CR3 = tan (60°) = 1.732 R3C = 4.59 * 10-3 1 We will choose a standard value of 120 nF for C, which yields R3 = 38.3 k. This is a standard value at 1% tolerance. Using these values, V2 will be V2 = V1 c V1 1 d = / -60° 1 + jCR3 2.0 11.33 From Eq. (11.33) we see that our inverter should also have a gain of 2 to restore the magnitude of V2 . The complete phase angle generator circuit is shown in Fig. 11.26c, where R4 = 10 k and R5 = 20 k have been chosen to produce the required gain. Now V3 is used to represent VCN . The additional unity gain buffer stage isolates the resistances associated with the inverter from the R-C phase generator. That way the inverter will not alter the phase angle. Finally, we must create the phase voltage VBN . Since the sum of the three-phase voltages is zero, we can write VBN = -VAN - VCN The simple op-amp summer in Fig. 11.26d will perform this mathematical operation. For the summer VBN = - c R8 R8 d VAN - c d VCN R6 R7 We require R6 = R7 = R8 . Since we are already using some 10-k resistors anyway, we just use three more here. The complete circuit is shown in Fig. 11.27 where one more unity gain buffer has been added at the potentiometer. This one isolates the R-C phase angle generator from the magnitude adjustment resistors. It may seem that we have used op-amps too liberally, requiring a total of four. However, most op-amp manufacturers package their op-amps in single (one op-amp), dual (two opamps), and quad (four op-amps) packages. Using a quad op-amp, we see that our circuit will require just one integrated circuit. As a final note, the op-amp power supply voltages must exceed the maximum input or output voltages at the op-amp terminals, which is 4 V. Therefore, we will specify +10 V supplies. R6 R1 ± – Rp R3 + ± – R2 V1 - VAN - 1 –––– jC The complete three-phase emulator with variable voltage magnitude. R8 R5 + + Figure 11.27 V2 - ± – R4 – ± R7 – ± + + V3=VCN - VBN - irwin11_541-576hr.qxd 570 • 6-08-2010 CHAPTER 11 15:11 Page 570 P O LY P H A S E C I R C U I T S SUMMARY ■ An important advantage of the balanced three-phase system is that it provides very smooth power delivery. ■ Because of the balanced condition, it is possible to analyze a circuit on a per-phase basis, thereby providing a significant computational shortcut to a solution. ■ A balanced three-phase voltage source has three sinusoidal voltages of the same magnitude and frequency, and each voltage is 120° out of phase with the others. A positivephase-sequence balanced voltage source is one in which Vbn lags Van by 120° and Vcn lags Vbn by 120°. ■ The relationships between wye- and delta-connected sources are shown in Table 11.1. ■ The three-phase terminology is shown in Table 11.3. ■ In a balanced system the voltages and currents sum to zero. TABLE 11.3 Three-phase terminology QUANTITY WYE DELTA Line current AIL B Phase current AIp B Ia , Ib , Ic Line-to-neutral voltage AVp B Van , Vbn , Vcn Phase voltage AVp B Vab , Vbc , Vca Line-to-line, phase-to-phase, line voltage AVLB Phase voltage AVp B Phase current AIp B Iab , Ibc , Ica 1. If the source/load connection is not wye–wye, then Van + Vbn + Vcn = 0 I a + I b + I c = 0 (no current in the neutral line) and transform the system to a wye–wye connection. 2. Determine the unknown phasors in the wye–wye connection and deal only with the phase a. 3. Convert the now-known phasors back to the corresponding phasors in the original connection. Vab + Vbc + Vca = 0 I ab + I bc + I ca = 0 ■ • The steps recommended for solving balanced three-phase ac circuits are as follows: ■ Power factor correction in a balanced three-phase environment is performed in the same manner as in the single-phase case. Three capacitors are put in parallel with the load to reduce the lagging phase caused by the three-phase load. PROBLEMS 11.1 Sketch a phasor representation of an abc-sequence 11.6 Find the equivalent Z of the network in Fig. P11.6. balanced three-phase Y-connected source, including Van , Vbn , and Vcn if Van = 120 / 15° V rms. 1⍀ –j1 ⍀ 2⍀ –j2 ⍀ 11.2 Sketch a phasor representation of a balanced three-phase system containing both phase voltages and line voltages if Van = 120 / 90° V rms. Label all magnitudes and assume an abc-phase sequence. 1⍀ 1⍀ 11.3 Sketch a phasor representation of a balanced three-phase system containing both phase voltages and line voltages if Van = 100 / 45° V rms. Label all magnitudes and assume an abc-phase sequence. j1 ⍀ Z 11.4 Sketch a phasor representation of a balanced three-phase 1⍀ system containing both phase voltages and line voltages if Vab = 208 / 60° V rms. Label all phasors and assume an abc-phase sequence. 11.5 A positive-sequence three-phase balanced wye voltage source has a phase voltage of Van = 240 / 90° V rms. Determine the line voltages of the source. 2⍀ –j1 ⍀ Figure P11.6 –j2 ⍀ irwin11_541-576hr.qxd 6-08-2010 15:11 Page 571 PROBLEMS 11.7 Find the equivalent impedances Z ab, Z bc, and Z ca in the network in Fig. P11.7. 11.11 A positive-sequence balanced three-phase wye-connected source supplies power to a balanced wye-connected load. The magnitude of the line voltages is 208 V rms. If the load impedance per phase is 36 + j12 , determine the line currents if / Van = 0°. a 1⍀ 11.12 In a three-phase balanced wye–wye system, the source is an abc-sequence set of voltages with Van = 120 / 60° rms. j1 ⍀ 1⍀ 571 The per phase impedance of the load is 12 + j16 . If the line impedance per phase is 0.8 + j1.4 , find the line currents and the voltages. 1⍀ j1 ⍀ j1 ⍀ c 11.13 An abc-sequence set of voltages feeds a balanced three-phase wye–wye system. If the line current in the a phase is 16.78. / 20.98° A rms, the line impedance is 1.2 + j1.8 , and the input voltage Vab = 440 / 70° V rms. find the load impedance. b Figure P11.7 11.14 An abc-phase-sequence three-phase balanced 11.8 Find the equivalent Z of the network in Fig. P11.8. 2⍀ – j1 ⍀ 1⍀ Z j1 ⍀ 11.15 An abc-phase-sequence three-phase balanced 1⍀ j1 ⍀ –j2 ⍀ –j2 ⍀ 1/2 ⍀ –j1 ⍀ –j2 ⍀ Figure P11.8 11.9 Find the equivalent Z of the network in Fig. P11.9. wye-connected 60-Hz source supplies a balanced deltaconnected load. The phase impedance in the load consists of a 20 resistor in series with a 50-mH inductor, and the phase voltage at the source is Van = 120 / 20° V rms V rms. If the line impedance is zero, find the line currents in the system. 11.16 An abc-phase-sequence three-phase balanced wyeconnected source supplies power to a balanced wye-connected load. The impedance per phase in the load is 14 + j12 . If the source voltage for the a phase is Van = 120 / 80° V rms, and the line impedance is zero, find the phase currents in the wye-connected source. 11.17 An abc-sequence balanced three-phase wye-connected 2⍀ a 1⍀ –j1 ⍀ Z wye-connected source supplies a balanced deltaconnected load. The impedance per phase in the delta load is 12 j9 . The line voltage at the source is Vab 120 23 / 40° V rms. If the line impedance is zero, find the line currents in the balanced wye–delta system. 1⍀ –j1 ⍀ 1⍀ c b 2⍀ 2⍀ – j1 ⍀ –j1 ⍀ d Figure P11.9 source supplies power to a balanced wye-connected load. The line impedance per phase is 1 + j5 , and the load impedance per phase is 25 j25 . If the source line voltage Vab 208 / 0° V rms find the line currents. 11.18 An abc-sequence balanced three-phase wye-connected source supplies power to a balanced wye-connected load. The line impedance per phase is 1 + j0 , and the load impedance per phase is 20 + j20 . If the source line voltage Vab is 100 / 0° V rms, find the line currents. 11.19 An abc-sequence set of voltages feeds a balanced 11.10 A positive-sequence balanced three-phase wye-connected source with a phase voltage of 120 V rms supplies power to a balanced wye-connected load. The per-phase load impedance is 40 + j10 . Determine the line currents in the circuit if / Van = 0° . three-phase wye–wye system. The line and load impedances are 1 + j1 and 10 + j10 , respectively. If the load voltage on the phase is VAN = 110 / 30° V rms, determine the line voltages of the input. irwin11_541-576hr.qxd 572 6-08-2010 CHAPTER 11 15:11 Page 572 P O LY P H A S E C I R C U I T S 11.20 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages. The load voltage on the a phase is VAN = 108.58 / 79.81° V rms, Zline = 1 + j1.4 , and Zload = 10 + j13 . Determine the input sequence of voltages. 11.21 A balanced abc-sequence of voltages feeds a balanced three-phase wye–wye system. The line and load impedance are 0.6 + j0.9 and 8 + j12 , respectively. The load voltage on the a phase is VAN = 116.63 / 10° V rms. Find the line voltage Vab . 11.22 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages. The load voltage on the a phase is VAN = 110 / 80° V rms, Z line = 1 + j1.4 , and Z load = 10 + j13 . Determine the input sequence of the line-to-neutral voltages. 11.30 In a balanced three-phase wye–wye system, the load impedance is 10 + j1 . The source has phase sequence abc and the line voltage Vab = 220 / 30° V rms. If the load voltage VAN = 120 / 0° V rms, determine the line impedance. 11.31 In a balanced three-phase wye–wye system, the load impedance is 20 + j12 . The source has an abc-phase sequence and Van = 120 / 0° V rms. If the load voltage is VAN = 111.49 /- 0.2° V rms, determine the magnitude of the line current if the load is suddenly short-circuited. 11.32 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages and Van = 120 / 40° V rms. If the a-phase line current and line impedance are known to be 7.10 / -10.28° A rms and 0.8 + j1 , respectively, find the load impedance. 11.33 In a balanced three-phase wye–wye system, the 11.23 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages. The load voltage on the a phase is VAN = 120 / 60° V rms, Z line = 2 + j1.4 , and Z load = 10 + j10 . Determine the input voltages. 11.24 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages. Z line = 1 + j1 , Z load = 14 + j12 , and the load voltage on the a phase is VAN = 440 / 30° V rms. Find the line voltage Vab. 11.25 A balanced abc-sequence of voltages feeds a balanced three-phase wye–wye system. The line and load impedance are 0.6 + j 0. and 8 + j12 respectively. The load voltage on the a phase is VAN = 116.63 / 10° V rms. Find the line voltage Vab . 11.26 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages. Zline = 1 + j1.8 , Zload = 14 + j12 , and the load voltage on the a phase is VAN = 398.1 / 17.99° V rms. Find the line voltage Vab. 11.27 In a balanced three-phase wye–delta system the source has an abc phase sequence and Van = 120 / 40° V rms. The line and load impedances are 0.5 + j0.4 and 24 + j18 , respectively. Find the delta currents in the load. 11.28 In a balanced three-phase wye–wye system, the load impedance is 8 + j4 . The source has phase sequence abc and Van = 120 / 0° V rms. If the load voltage is VAN = 111.62 / -1.33° V rms, determine the line impedance. 11.29 In a balanced three-phase wye–wye system, the total power loss in the lines is 400 W. VAN = 105.28 / 31.56° V rms and the power factor of the load is 0.77 lagging. If the line impedance is 2 + j1 , determine the load impedance. source is an abc-sequence set of voltages and Van = 120 / 50° V rms. The load voltage on the a phase is 110.65 / 29.03° V rms and the load impedance is 16 + j20 . Find the line impedance. 11.34 An abc-sequence set of voltages feeds a balanced three-phase wye–wye system. If the input voltage Van = 440 / 10° V rms, the load voltage on the a phase is VAN = 398.32 / 8.72° V rms, and Zload is 20 + j24 , find the line impedance. 11.35 An abc-phase-sequence balanced three-phase source feeds a balanced load. The system is wye–wye connected. The load impedance is 10 + j6 , the line impedance is 1 + j0.5 V, and / VAN = 60°. The total power loss in the lines is 470.44 W. Find VAN and the magnitude of the source voltage. 11.36 An abc-sequence balanced three-phase source feeds a balanced wye–wye system. Zline = 0.8 + j0.2 , Zload = 12 + j6 , / VAN = 30°. The total power absorbed by the load is 2 kW. Determine the total power loss in the lines. 11.37 An abc-phase-sequence three-phase balanced wye-connected 60-Hz source supplies a balanced delta-connected load. The phase impedance in the load consists of a 20- resistor series with a 20-mH inductor, and the phase voltage at the source is Van = 120 / 30° V rms. If the line impedance is zero, find the line currents in the system. 11.38 In a three-phase balanced system, a delta-connected source supplies power to a wye-connected load. If the line impedance is 0.2 + j 0.4 , the load impedance 3 + j2 , and the source phase voltage Vab = 208 / 10° V rms, find the magnitude of the line voltage at the load. 11.39 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages and Van = 120 / 50° V rms. The load voltage on the a phase is 110 / 50° V rms, and the load impedance is 16 j20 . Find the line impedance. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 573 PROBLEMS 11.40 In a balanced three-phase wye–wye system, the source is an abc-sequence set of voltages and Van 120 / 40° V rms. If the a-phase line current and line impedance are known to be 6 / 15° A rms and 1 + j1 , respectively, find the load impedance. 11.41 An abc-phase-sequence three-phase balanced wye-connected source supplies a balanced deltaconnected load. The impedance per phase in the delta load is 12 + j6 . The line voltage at the source is Vab 120 23 / 40° V rms. If the line impedance is zero, find the line currents in the balanced wye–delta system. 11.42 In a balanced three-phase delta–wye system, the source has an abc-phase sequence. The line and load impedances are 0.6 + j 0.3 and 12 + j7 , respectively. If the line current I aA 9.6 /- 20° A rms, determine the phase voltages of the source. 11.43 In a three-phase balanced system, a delta-connected source supplies power to a wye-connected load. If the line impedance is 0.2 + j0.4 , the load impedance 6 + j4 , and the source phase voltage Vab = 210 / 40° V rms, find the magnitude of the line voltage at the load. 11.44 An abc-sequence set of voltages feeds a balanced threephase wye–wye system. If Van = 440 / 30° V rms, VAN = 413.28 / 29.78° V rms, and Z line = 2 + j1.5 , find the load impedance. 11.45 An abc-phase-sequence three-phase balanced wyeconnected source supplies a balanced delta-connected load. The impedance per phase of the delta load is 20 + j4 . If VAB = 115 / 35° V rms, find the line current. 11.46 An abc-phase-sequence three-phase balanced wyeconnected source supplies power to a balanced deltaconnected load. The impedance per phase in the load is 14 + j7 . If the source voltage for the a phase is Van = 120 / 80° V rms and the line impedance is zero, find the phase currents in the wye-connected source. 11.47 In a three-phase balanced delta–delta system, the source has an abc-phase sequence. The line and load impedances are 0.3 + j0.2 and 9 + j6 , respectively. If the load current in the delta is I AB = 15 / 40° A rms, find the phase voltages of the source. 11.48 An abc-phase-sequence three-phase balanced wyeconnected source supplies a balanced delta-connected load. The impedance per phase of the delta load is 10 + j8 . If the line impedance is zero and the line current in the a phase is known to be I aA = 28.10 / -28.66° A rms, find the load voltage VAB . 11.49 In a balanced three-phase wye–delta system, the source has an abc-phase sequence and Van = 120 / 0° V rms. If the line impedance is zero and the line current I aA = 5 / 20° A rms , find the load impedance per phase in the delta. 573 11.50 A three-phase load impedance consists of a balanced wye in parallel with a balanced delta. What is the equivalent wye load and what is the equivalent delta load if the phase impedances of the wye and delta are 6 + j3 and 15 + j10 , respectively? 11.51 In a balanced three-phase system, the abc-phasesequence source is wye connected and Van = 120 / 20° V rms. The load consists of two balanced wyes with phase impedances of 8 + j2 and 12 + j3 . If the line impedance is zero, find the line currents and the phase current in each load. 11.52 In a balanced three-phase delta–delta system, the source has an abc-phase sequence. The phase angle for the source voltage is / Vab = 40° and I ab = 4 / 15° A rms. If the total power absorbed by the load is 1400 W, find the load impedance. 11.53 In a balanced three-phase system, the source is a balanced wye with an abc-phase sequence and Vab = 215 / 50° V rms. The load is a balanced wye in parallel with a balanced delta. The phase impedance of the wye is 5 + j3 , and the phase impedance of the delta is 18 + j12 . If the line impedance is 1 + j0.8 , find the line currents and the phase currents in the loads. 11.54 In a balanced three-phase system, the source is a balanced wye with an abc-phase sequence and Vab = 208 / 60° V rms. The load consists of a balanced wye with a phase impedance of 8 + j5 in parallel with a balanced delta with a phase impedance of 21 + j12 . If the line impedance is 1.2 + j1 , find the phase currents in the balanced wye load. 11.55 In a balanced three-phase system, the source has an abc-phase sequence and is connected in delta. There are two loads connected in parallel. The line connecting the source to the loads has an impedance of 0.2 + j0.1 . Load 1 is connected in wye, and the phase impedance is 4 + j2 . Load 2 is connected in delta, and the phase impedance is 12 + j9 . The current I AB in the delta load is 16 / 45°A rms. Find the phase voltage of the source. 11.56 In a balanced three-phase system, the source has an abc-phase sequence and is connected in delta. There are two parallel wye-connected loads. The phase impedance of load 1 and load 2 is 4 + j4 and 10 + j4 , respectively. The line impedance connecting the source to the loads is 0.3 + j0.2 . If the current in the a phase of load 1 is I AN1 = 10 / 20° A rms, find the delta currents in the source. 11.57 An abc-phase-sequence balanced three-phase source feeds a balanced load. The system is connected wyewye and / Van = 0°. The line impedance is 0.5 + j0.2 , the load impedance is 16 + j10 , and the total power absorbed by the load is 2000 W. Determine the magnitude of the source voltage Van. irwin11_541-576hr.qxd 574 6-08-2010 CHAPTER 11 15:11 Page 574 P O LY P H A S E C I R C U I T S 11.58 A balanced three-phase delta-connected source supplies power to a load consisting of a balanced delta in parallel with a balanced wye. The phase impedance of the delta is 24 + j12 , and the phase impedance of the wye is 12 + j8 . The abc-phase-sequence source voltages are Vab = 440 / 60° V rms, Vbc = 440 /- 60° V rms, and Vca = 440 /- 180° V rms, and the line impedance per phase is 1 + j0.08 . Find the line currents and the power absorbed by the wye-connected load. 11.59 The magnitude of the complex power (apparent power) supplied by a three-phase balanced wye–wye system is 3600 VA. The line voltage is 208 V rms. If the line impedance is negligible and the power factor angle of the load is 25°, determine the load impedance. 11.60 An abc-sequence wye-connected source having a phase-a voltage of 120 / 0° V rms is attached to a wye-connected load having a per-phase impedance of 100 / 70° . If the line impedance is 1 / 20° Æ , determine the total complex power produced by the voltage sources and the real and reactive power dissipated by the load. 11.61 A three-phase balanced wye–wye system has a line voltage of 208 V rms. The line current is 6 A rms and the total real power absorbed by the load is 1800 W. Determine the load impedance per-phase, if the line impedance is negligible. 11.62 A three-phase abc-sequence wye-connected source supplies 14 kVA with a power factor of 0.75 lagging to a delta load. If the delta load consumes 12 kVA at a power factor of 0.7 lagging and has a phase current of 10 /- 30° A rms, determine the per-phase impedance of the load and the line. 11.63 A balanced three-phase source serves the following loads: Load 1: 60 kVA at 0.8 pf lagging Load 2: 30 kVA at 0.75 pf lagging The line voltage at the load is 208 V rms at 60 Hz. Determine the line current and the combined power factor at the load. 11.64 A balanced three-phase source serves two loads: Load 1: 36 kVA at 0.8 pf lagging Load 2: 18 kVA at 0.6 pf lagging The line voltage at the load is 208 V rms at 60 Hz. Find the line current and the combined power factor at the load. 11.65 The following loads are served by a balanced three-phase source: Load 1: 18 kVA at 0.8 pf lagging Load 2: 8 kVA at 0.8 pf leading Load 3: 12 kVA at 0.75 pf lagging The load voltage is 208 V rms at 60 Hz. If the line impedance is negligible, find the power factor at the source. 11.66 A balanced three-phase source serves the following loads: Load 1: 20 kVA at 0.8 pf lagging Load 2: 10 kVA at 0.7 pf leading Load 3: 10 kW at unity pf Load 4: 16 kVA at 0.6 pf lagging The line voltage at the load is 208 V rms at 60 Hz, and the line impedance is 0.02 + j0.04 . Find the line voltage and power factor at the source. 11.67 A small shopping center contains three stores that represent three balanced three-phase loads. The power lines to the shopping center represent a three-phase source with a line voltage of 13.8 kV rms. The three loads are Load 1: 400 kVA at 0.9 pf lagging Load 2: 200 kVA at 0.85 pf lagging Load 3: 100 kVA at 0.90 pf lagging Find the power line current. 11.68 The following loads are served by a balanced three-phase source: Load 1: 20 kVA at 0.8 pf lagging Load 2: 4 kVA at 0.8 pf leading Load 3: 10 kVA at 0.75 pf lagging The load voltage is 208 V rms at 60 Hz. If the line impedance is negligible, find the power factor at the source. 11.69 A balanced three-phase source supplies power to three loads: Load 1: 30 kVA at 0.8 pf lagging Load 2: 24 kW at 0.6 pf leading Load 3: unknown If the line voltage and total complex power at the load are 208 V rms and 60 / 0° kVA, respectively, find the unknown load. 11.70 A balanced three-phase source supplies power to three loads. The loads are Load 1: 24 kVA at 0.6 pf lagging Load 2: 10 kW at 0.75 pf lagging Load 3: unknown If the line voltage at the load is 208 V rms, the magnitude of the total complex power is 35.52 kVA, and the combined power factor at the load is 0.88 lagging, find the unknown load. 11.71 A balanced three-phase source supplies power to three loads: Load 1: 18 kVA at 0.8 pf lagging Load 2: 10 kW at 0.6 pf leading Load 3: unknown If the line voltage at the loads is 208 V rms, the line current at the source is 116.39 A rms, and the combined power factor at the load is 0.86 lagging, find the unknown load. irwin11_541-576hr.qxd 6-08-2010 15:11 Page 575 PROBLEMS 11.72 A balanced three-phase source supplies power to three loads: 575 11.77 Find C in the network in Fig. P11.77 such that the total load has a power factor of 0.9 lagging. Load 1: 30 kVA at 0.8 pf lagging Load 2: 24 kW at 0.6 pf leading + Load 3: unknown The line voltage at the load and line current at the source are 208 V rms and 166.8 A rms, respectively. If the combined power factor at the load is unity, find the unknown load. C 4.6 kV rms Balanced three-phase source 60 Hz - Balanced three-phase load 6 MVA 0.8 pf lagging C C 11.73 A balanced three-phase source supplies power to three loads: Load 1: 24 kW at 0.8 pf lagging Load 2: 10 kVA at 0.7 pf leading Figure P11.77 Load 3: unknown If the line voltage at the load is 208 V rms, the magnitude of the total complex power is 41.93 kVA, and the combined power factor at the load is 0.86 lagging, find the unknown load. 11.78 Find the value of C in Fig. P11.78 such that the total load has a power factor of 0.87 lagging. 11.74 A three-phase abc-sequence wye-connected source with Van = 220 / 0° V rms supplies power to a wye-connected load that consumes 50 kW of power in each phase at a pf of 0.8 lagging. Three capacitors are found that each have an impedance of -j2.0 , and they are connected in parallel with the load in a wye configuration. Determine the power factor of the combined load as seen by the source. + 34.5 kV rms Balanced three-phase source 60 Hz C C - C Balanced three-phase load 20 MVA 0.707 pf lagging Figure P11.78 11.75 If the three capacitors in the network in Problem 11.74 are connected in a delta configuration, determine the power factor of the combined load as seen by the source. 11.79 Find C in the network in Fig. P11.79 so that the total load has a power factor of 0.9 leading. 11.76 Find C in the network in Fig. P11.76 such that the total load has a power factor of 0.87 leading. + C 4.6 kV rms + 34.5 kV rms Balanced three-phase source 60 Hz Figure P11.76 C - C C Balanced three-phase load 20 MVA 0.707 pf lagging Balanced three-phase source 60 Hz - Balanced three-phase load 6 MVA 0.8 pf lagging C C Figure P11.79 irwin11_541-576hr.qxd 576 6-08-2010 15:11 CHAPTER 11 Page 576 P O LY P H A S E C I R C U I T S 11.80 A standard practice for utility companies is to divide customers into single-phase users and three-phase users. The utility must provide three-phase users, typically industries, with all three phases. However, single-phase users, residential and light commercial, are connected to only one phase. To reduce cable costs, all single-phase users in a neighborhood are connected together. This means that even if the three-phase users present perfectly balanced loads to the power grid, the single-phase loads will never be in balance, resulting in current flow in the neutral connection. Consider the 60-Hz, abc-sequence network in Fig. P11.80. With a line voltage of 416 / 30° V rms, phase a supplies the single-phase users on A Street, phase b supplies B Street, and phase c supplies C Street. Furthermore, the three-phase industrial load, which is connected in delta, is balanced. Find the neutral current. A a 240 0° V rms ± – Three-phase 36 kW pf=0.5 lagging B b 240 –120° V rms ± – C c IAN ± – 240 120° V rms n A Street 48 kW pf=1 InN ICN IBN B Street 30 kW pf=1 C Street 60 kW pf=1 N Figure P11.80 TYPICAL PROBLEMS FOUND ON THE FE EXAM • 11PFE-1 A wye-connected load consists of a series RL impedance. Measurements indicate that the rms voltage across each element is 84.85 V. If the rms line current is 6A, find the total complex power for the three-phase load configuration. a. 1.25 / -45° kVA b. 4.32 / 30° kVA c. 3.74 / 60° kVA d. 2.16 / 45° kVA 11PFE-2 A balanced three-phase delta-connected load consists of an impedance of 12 + j12 . If the line voltage at the load is measured to be 230 V rms, find the total real power absorbed by the three-phase configuration. a. 6.62 kW b. 2.42 kW c. 3.36 kW d. 5.82 kW 11PFE-3 Two balanced three-phase loads are connected in parallel. One load with a phase impedance of 24 + j18 is connected in delta, and the other load has a phase impedance of 6 + j4 and is connected in wye. If the line-to-line voltage is 208 V rms, determine the line current. a. 15.84 / -60.25°A rms b. 28.63 / -35.02°A rms c. 40.49 / 30.27°A rms d. 35.32 / 90.53°A rms 11PFE-4 The total complex power at the load of a three-phase balanced system is 24 / 30° kVA. Find the real power per phase. a. 3.24 kW b. 4.01 kW c. 6.93 kW d. 8.25 kW 11PFE-5 A balanced three-phase load operates at 90 kW with a line voltage at the load of 480 / 0° V rms at 60 Hz. The apparent power of the three-phase load is 100 kVA. It is known that the load has a lagging power factor. What is the total three-phase reactive power of the load? a. 22.43 kvar b. 30.51 kvar c. 25.35 kvar d. 43.59 kvar irwin12_577-666hr.qxd 6-08-2010 15:14 Page 577 12 C H A PT E R VARIABLE-FREQUENCY NETWORK PERFORMANCE THE LEARNING GOALS FOR THIS CHAPTER ARE: ■ Understand the variable-frequency performance of the basic circuit elements: R, L, and C ■ Learn the different types of network functions and the definition of poles and zeros ■ Be able to sketch a Bode plot for a network function ■ Know how to create a Bode plot ■ Know how to analyze series and parallel resonant circuits ■ Be introduced to the concepts of magnitude and frequency scaling ■ Learn the characteristics of basic filters such low-pass, high-pass, band-pass, and band rejection ■ Know how to analyze basic passive and active filters David Redfern/Redferns/Getty Images, Inc. M Music and Frequency Time and frequency are two sides of Your hearing must have a bandwidth that extends to fre- the same coin, describing signals in two ways. We hear music, quencies at least as high as the highest frequency present in for example, as a time signal but the sampling rate for its digi- the musical signal. Voice or other signals can be modulated to tal recording, mastering, and manufacturing depends on signal a band of much higher frequencies, so high in some cases that bandwidth—a frequency description. We utilize both “lan- they can be heard only by dogs or Superman. High-frequency guages”—time and frequency—to design recording and play- hearing tends to deteriorate as you get older. This can be an back equipment. For iPods and MP3 players on laptops, the advantage when you’re young. For example, you can download bandwidth has to be large enough to reproduce music in fre- high-frequency ringtones that are inaudible to the average pro- quency ranges from bass to violin and from tuba to flute. For fessor. On the other hand, some convenience stores discour- reference, concert flutes have frequencies from 262 Hz to over age loitering by broadcasting high-frequency signals that most 2 kHz (middle C to three octaves above). adults can’t hear, but that are very irritating to teenagers. 577 irwin12_577-666hr.qxd 578 6-08-2010 CHAPTER 12 15:14 Page 578 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E Variable frequency is as important in electric circuits as it is desired frequency spectra—low-pass, high-pass, or band- in describing musical signals. Frequency response plots of pass—for recording, communication, and radar systems. Your voltage transfer functions show magnitude in decibels and cell phone, television, and iPod or MP3 player use these phase in degrees on linear scales versus radian frequency on principles. The music beat delivered from high quality record- a logarithmic scale. Filter design using variable frequency ings and the clear tones of live performances both uplift the techniques results in series or parallel circuits that yield listener—provided the bandwidth is adequate. 12.1 In previous chapters we investigated the response of RLC networks to sinusoidal inputs. In particular, we considered 60-Hz sinusoidal inputs. In this chapter we allow the frequency of excitation to become a variable and evaluate network performance as a function of frequency. To begin, let us consider the effect of varying frequency on elements with which we are already quite familiar—the resistor, inductor, and capacitor. The frequency-domain impedance of the resistor shown in Fig. 12.1a is Variable FrequencyResponse Analysis Z R = R = R / 0° The magnitude and phase are constant and independent of frequency. Sketches of the magnitude and phase of ZR are shown in Figs. 12.1b and c. Obviously, this is a very simple situation. For the inductor in Fig. 12.2a, the frequency-domain impedance ZL is Z L = jL = L / 90° Figure 12.1 Frequency-independent impedance of a resistor. ZR R Magnitude of ZR () (a) R 0 Frequency Phase of ZR (degrees) (b) 0 0 Frequency (c) irwin12_577-666hr.qxd 6-08-2010 15:14 Page 579 SECTION 12.1 V A R I A B L E F R E Q U E N C Y- R E S P O N S E A N A LY S I S 579 Figure 12.2 ZL L Magnitude of ZL () (a) 0 0 Frequency (b) Phase of ZL (degrees) ±90° 0 0 Frequency (c) The phase is constant at 90°, but the magnitude of ZL is directly proportional to frequency. Figs. 12.2b and c show sketches of the magnitude and phase of ZL versus frequency. Note that at low frequencies the inductor’s impedance is quite small. In fact, at dc, ZL is zero, and the inductor appears as a short circuit. Conversely, as frequency increases, the impedance also increases. Next consider the capacitor of Fig. 12.3a. The impedance is ZC = 1 1 = / -90° jC C Once again the phase of the impedance is constant, but now the magnitude is inversely proportional to frequency, as shown in Figs. 12.3b and c. Note that the impedance approaches infinity, or an open circuit, as approaches zero and ZC approaches zero as approaches infinity. Now let us investigate a more complex circuit: the RLC series network in Fig. 12.4a. The equivalent impedance is Z eq = R + jL + 1 jC or Z eq = (j)2LC + jRC + 1 jC Frequency-dependent impedance of an inductor. irwin12_577-666hr.qxd 580 6-08-2010 CHAPTER 12 15:14 Page 580 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E Figure 12.3 Frequency-dependent impedance of a capacitor. ZC C Magnitude of ZC () (a) R 0 0 Frequency Phase of ZC (degrees) (b) 0 –90° 0 Frequency (c) Sketches of the magnitude and phase of this function are shown in Figs. 12.4b and c. Note that at very low frequencies, the capacitor appears as an open circuit, and, therefore, the impedance is very large in this range. At high frequencies, the capacitor has very little effect and the impedance is dominated by the inductor, whose impedance keeps rising with frequency. As the circuits become more complicated, the equations become more cumbersome. In an attempt to simplify them, let us make the substitution j=s. (This substitution has a more important meaning, which we will describe in later chapters.) With this substitution, the expression for Zeq becomes Z eq = s2LC + sRC + 1 sC If we review the four circuits we investigated thus far, we will find that in every case the impedance is the ratio of two polynomials in s and is of the general form Z(s) = N(s) = D(s) am sm + am - 1 sm - 1 + p + a1 s + a0 bn s n + bn - 1 sn - 1 + p + b1 s + b0 12.1 where N(s) and D(s) are polynomials of order m and n, respectively. An extremely important aspect of Eq. (12.1) is that it holds not only for impedances but also for all voltages, currents, admittances, and gains in the network. The only restriction is that the values of all circuit elements (resistors, capacitors, inductors, and dependent sources) must be real numbers. irwin12_577-666hr.qxd 6-08-2010 15:14 Page 581 SECTION 12.1 V A R I A B L E F R E Q U E N C Y- R E S P O N S E A N A LY S I S Figure 12.4 L R 581 Zeq Frequency-dependent impedance of an RLC series network. C Magnitude of Zeq () (a) R 0 1 0 = ——– 1LC Frequency Phase of Zeq (degrees) (b) ±90° 0 –90° 0 Frequency (c) Let us now demonstrate the manner in which the voltage across an element in a series RLC network varies with frequency. Consider the network in Fig. 12.5a. We wish to determine the variation of the output voltage as a function of frequency over the range from 0 to 1 kHz. Using voltage division, we can express the output as Vo = ° R R + jL + 1 ¢ jC VS jCR 2 (j) LC + jCR + 1 b VS Using the element values, we find that the equation becomes Vo = a (j)A37.95 * 10-3 B 12.1 SOLUTION or, equivalently, Vo = a EXAMPLE (j)2 A2.53 * 10-4 B + jA37.95 * 10-3 B + 1 b 10 / 0° irwin12_577-666hr.qxd CHAPTER 12 15:14 Page 582 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E Figure 12.5 L=0.1 H (a) Network and (b) its frequency-response simulation. C=2.53 mF VS=10 0° V ± – R=15 + Vo - (a) Magnitude 101 100 10–1 10–2 100 101 Frequency (rad) 102 100 101 Frequency (rad) 102 100 Phase (deg) 582 6-08-2010 50 0 –50 –100 (b) The resultant magnitude and phase characteristics are semilog plots in which the frequency is displayed on the log axis. The plots for the function Vo are shown in Fig. 12.5b. In subsequent sections we will illustrate that the use of a semilog plot is a very useful tool in deriving frequency-response information. As an introductory application of variable frequency-response analysis and characterization, let us consider a stereo amplifier. In particular, we should consider first the frequency range over which the amplifier must perform and then exactly what kind of performance we desire. The frequency range of the amplifier must exceed that of the human ear, which is roughly 50 Hz to 15,000 Hz. Accordingly, typical stereo amplifiers are designed to operate in the frequency range from 50 Hz to 20,000 Hz. Furthermore, we want to preserve the fidelity of the signal as it passes through the amplifier. Thus, the output signal should be an exact duplicate of the input signal times a gain factor. This requires that the gain be independent of frequency over the specified frequency range of 50 Hz to 20,000 Hz. An ideal sketch of this requirement for a gain of 1000 is shown in Fig. 12.6, where the midband region is defined as irwin12_577-666hr.qxd 6-08-2010 15:14 Page 583 SECTION 12.1 V A R I A B L E F R E Q U E N C Y- R E S P O N S E A N A LY S I S 583 Figure 12.6 Amplifier frequency-response requirements. Gain (*1000) 1.0 0.8 0.6 0.4 0.2 fLO 0 1 10 fHI 100 1k 10 k 100 k 1M Frequency (Hz) Ro Cin Figure 12.7 + + ± – vin(t) Rin vS(t) + 1000vin(t) Co - - - Ro=100 Co=79.58 nF Cin=3.18 nF Rin=1 M vo(t) (a) Ro 1/sCin + + + Rin VS(s) - ± – Vin(s) 1000Vin(s) 1/sCo Vo(s) - (b) that portion of the plot where the gain is constant and is bounded by two points, which we will refer to as fLO and fHI . Notice once again that the frequency axis is a log axis and, thus, the frequency response is displayed on a semilog plot. A model for the amplifier described graphically in Fig. 12.6 is shown in Fig. 12.7a, with the frequency-domain equivalent circuit in Fig. 12.7b. If the input is a steady-state sinusoid, we can use frequency-domain analysis to find the gain Gv(j) = Vo(j) VS(j) which with the substitution s=j can be expressed as Gv(s) = Vo(s) VS(s) Using voltage division, we find that the gain is Gv(s) = Vo(s) VS(s) Vin(s) Vo(s) = VS(s) Vin(s) = c Rin 1兾sCo d (1000) c d Rin + 1兾sCin Ro + 1兾sCo or Gv(s) = c sCin Rin 1 d (1000) c d 1 + sCin Rin 1 + sCo Ro Amplifier equivalent network. irwin12_577-666hr.qxd CHAPTER 12 15:14 Page 584 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E Figure 12.8 1.0 Exact and approximate amplifier gain versus frequency plots. 0.8 Gain (*1000) 584 6-08-2010 0.6 0.4 0.2 fLO 0 1 100 10 fHI 1k 10 k 100 k Frequency (Hz) 1M Using the element values in Fig. 12.7a, Gv(s) = c 40,000 s d (1000) c d s + 100 s + 40,000 where 100 and 40,000 are the radian equivalents of 50 Hz and 20,000 Hz, respectively. Since s=j, the network function is indeed complex. An exact plot of Gv(s) is shown in Fig. 12.8 superimposed over the sketch of Fig. 12.6. The exact plot exhibits smooth transitions at fLO and fHI ; otherwise the plots match fairly well. Let us examine our expression for Gv(s) more closely with respect to the plot in Fig. 12.8. Assume that f is well within the midband frequency range; that is, fLO V f V fHI or 100 V ∑s∑ V 40,000 Under these conditions, the network function becomes s 1 d Gv(s) L c d (1000) c s 1 + 0 or Gv(s) = 1000 Thus, well within midband, the gain is constant. However, if the frequency of excitation decreases toward fLO , then ∑s∑ is comparable to 100 and Gv(s) L c s d (1000) s + 100 Since Rin Cin =1/100, we see that Cin causes the rolloff in gain at low frequencies. Similarly, when the frequency approaches fHI , the gain rolloff is due to Co . Through this amplifier example, we have introduced the concept of frequency-dependent networks and have demonstrated that frequency-dependent network performance is caused by the reactive elements in a network. N E T W O R K F U N C T I O N S In the previous section, we introduced the term voltage gain, Gv(s). This term is actually only one of several network functions, designated generally as H(s), which define the ratio of response to input. Since the function describes a reaction due to an excitation at some other point in the circuit, network functions are also called transfer functions. Furthermore, transfer functions are not limited to voltage ratios. Since in electrical networks inputs and outputs can be either voltages or currents, there are four possible network functions, as listed in Table 12.1. There are also driving point functions, which are impedances or admittances defined at a single pair of terminals. For example, the input impedance of a network is a driving point function. irwin12_577-666hr.qxd 6-08-2010 15:14 Page 585 SECTION 12.1 TABLE 12.1 V A R I A B L E F R E Q U E N C Y- R E S P O N S E A N A LY S I S 585 Network transfer functions INPUT OUTPUT TRANSFER FUNCTION Voltage Current Current Voltage Voltage Voltage Current Current Voltage gain Transimpedance Current gain Transadmittance SYMBOL Gv(s) Z(s) Gi(s) Y(s) We wish to determine the transfer admittance CI2(s)/V1(s)D and the voltage gain of the network shown in Fig. 12.9. EXAMPLE 12.2 The mesh equations for the network are AR1 + sLBI 1(s) - sLI 2(s) = V1(s) -sLI 1(s) + a R2 + sL + SOLUTION 1 b I (s) = 0 sC 2 V2(s) = I 2(s)R2 Solving the equations for I2(s) yields I 2(s) = sLV1(s) AR1 + sLB AR2 + sL + 1兾sCB - s2L2 Therefore, the transfer admittance CI2(s)/V1(s)D is I 2(s) YT(s) = V1(s) = LCs2 AR1 + R2 BLCs2 + AL + R1 R2 CBs + R1 and the voltage gain is Gv(s) = V2(s) V1(s) = LCR2 s2 AR1 + R2 BLCs2 + AL + R1 R2 CBs + R1 R1 Figure 12.9 1 — sC Circuit employed in Example 12.2. + V1(s) ± – I1(s) sL I2(s) V2(s) R2 - POLES AND ZEROS As we have indicated, the network function can be expressed as the ratio of the two polynomials in s. In addition, we note that since the values of our circuit elements, or controlled sources, are real numbers, the coefficients of the two polynomials will be real. Therefore, we will express a network function in the form H(s) = N(s) = D(s) am sm + am - 1 sm - 1 + p + a1 s + a0 bn sn + bn - 1 sn - 1 + p + b1 s + b0 12.2 where N(s) is the numerator polynomial of degree m and D(s) is the denominator polynomial of degree n. Equation (12.2) can also be written in the form H(s) = K0 As - z1 BAs - z2 B p As - zm B As - p1 B As - p2 B p As - pn B 12.3 irwin12_577-666hr.qxd 586 6-08-2010 CHAPTER 12 15:14 Page 586 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E where K0 is a constant, z1 , p , zm are the roots of N(s), and p1 , p , pn are the roots of D(s). Note that if s=z1 , or z2 , p , zm , then H(s) becomes zero and hence z1 , p , zm are called zeros of the transfer function. Similarly, if s=p1 , or p2 , p , pn , then H(s) becomes infinite and, therefore, p1 , p , pn are called poles of the function. The zeros or poles may actually be complex. However, if they are complex, they must occur in conjugate pairs since the coefficients of the polynomial are real. The representation of the network function specified in Eq. (12.3) is extremely important and is generally employed to represent any linear time-invariant system. The importance of this form lies in the fact that the dynamic properties of a system can be gleaned from an examination of the system poles. Learning Assessments E12.1 Find the driving-point impedance at VS(s) in the amplifier shown in Fig. 12.7b. ANSWER: 1 sCin Z(s) = Rin + = c1 + a E12.2 Find the pole and zero locations in hertz and the value of K0 for the amplifier network in Fig. 12.7. 100 b d M. s ANSWER: z1=0 Hz (dc); p1=–50 Hz; p2=–20,000 Hz; K0=A4 * 107 B . E12.3 Determine the voltage transfer function Vo (s)/Vi (s) as a function of s in Fig. PE12.3. R1 + Vi(s) – 12.2 Sinusoidal Frequency Analysis s ANSWER: 1 sC1 R1C2 s 2+ C1R2 + C2R2 + C1R1 R1R2C1C2 s+ 1 R1R2C1C2 + R2 1 Vo(s) sC2 – Figure E12.3 Although in specific cases a network operates at only one frequency (e.g., a power system network), in general we are interested in the behavior of a network as a function of frequency. In a sinusoidal steady-state analysis, the network function can be expressed as H(j) = M()ej() 12.4 where M() = ∑H(j)∑and () is the phase. A plot of these two functions, which are commonly called the magnitude and phase characteristics, displays the manner in which the response varies with the input frequency . We will now illustrate the manner in which to perform a frequency-domain analysis by simply evaluating the function at various frequencies within the range of interest. F R EQ U E N C Y R E S P O N S E U S I N G A B O D E P LOT If the network characteristics are plotted on a semilog scale (that is, a linear scale for the ordinate and a logarithmic scale for the abscissa), they are known as Bode plots (named after Hendrik W. Bode). This graph is a irwin12_577-666hr.qxd 6-08-2010 15:14 Page 587 SECTION 12.2 S I N U S O I D A L F R E Q U E N C Y A N A LY S I S powerful tool in both the analysis and design of frequency-dependent systems and networks such as filters, tuners, and amplifiers. In using the graph, we plot 20 log 10 M() versus log 10 () instead of M() versus . The advantage of this technique is that rather than plotting the characteristic point by point, we can employ straight-line approximations to obtain the characteristic very efficiently. The ordinate for the magnitude plot is the decibel (dB). This unit was originally employed to measure the ratio of powers; that is, number of dB = 10 log 10 P2 P1 12.5 If the powers are absorbed by two equal resistors, then @V2 @ 兾R 2 number of dB = 10 log 10 = 20 log 10 @V1 @ 兾R 2 @V2 @ @V1 @ = 10 log 10 = 20 log 10 @I 2 @ 2R @I 1 @ 2R @I 2 @ 12.6 @I 1 @ The term dB has become so popular that it is now used for voltage and current ratios, as illustrated in Eq. (12.6), without regard to the impedance employed in each case. In the sinusoidal steady-state case, H(j) in Eq. (12.3) can be expressed in general as H(j) = 2 K0(j);N A1 + j1 B C 1 + 23 Aj3 B + Aj3 B D p 2 A1 + ja B C 1 + 2b Ajb B + Ajb B D p 12.7 Note that this equation contains the following typical factors: 1. A frequency-independent factor K0>0 2. Poles or zeros at the origin of the form j; that is, (j)+N for zeros and (j)-N for poles 3. Poles or zeros of the form (1 + j) 4. Quadratic poles or zeros of the form 1 + 2(j) + (j)2 Taking the logarithm of the magnitude of the function H(j) in Eq. (12.7) yields 20 log 10∑H(j)∑ = 20 log 10 K0 ; 20N log 10∑j∑ + 20 log 10 @1 + j1 @ + 20 log 10 @1 + 23 Aj3 B + Aj3 B @ 2 12.8 + p - 20 log 10 @1 + ja @ 2 - 20 log 10 @1 + 2b Ajb B + Ajb B @ p Note that we have used the fact that the log of the product of two or more terms is equal to the sum of the logs of the individual terms, the log of the quotient of two terms is equal to the difference of the logs of the individual terms, and log 10 An = n log 10 A. The phase angle for H(j) is / H(j) = 0 ; N(90°) + tan-1 1 + tan-1 a 23 3 1 - 223 b 2b b p + p - tan-1 a - tan-1 a b 1 - 22b 12.9 587 Page 588 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E Magnitude characteristic Figure 12.10 Log magnitude gain (dB) Phase characteristic 0.1 1.0 10 0 100 (rad/s:log scale) (a) Log magnitude gain (dB) Magnitude characteristic with slope of –20N dB/decade 0 Phase characteristic –N(90°) 1.0 (rad/s:log scale) (b) Log magnitude gain (dB) Magnitude and phase characteristics for a constant term and poles and zeros at the origin. 20 log10K0 Phase (deg) CHAPTER 12 15:14 Phase (deg) 588 6-08-2010 0 Phase characteristic Magnitude characteristic with slope of ±20N dB/decade 1.0 (rad/s:log scale) ±N(90°) Phase (deg) irwin12_577-666hr.qxd (c) As Eqs. (12.8) and (12.9) indicate, we will simply plot each factor individually on a common graph and then sum them algebraically to obtain the total characteristic. Let us examine some of the individual terms and illustrate an efficient manner in which to plot them on the Bode diagram. Constant Term The term 20 log 10 K0 represents a constant magnitude with zero phase shift, as shown in Fig. 12.10a. Poles or Zeros at the Origin Poles or zeros at the origin are of the form (j);N, where + is used for a zero and-is used for a pole. The magnitude of this function is ;20N log 10 , which is a straight line on semilog paper with a slope of ;20N dB兾decade; that is, the value will change by 20N each time the frequency is multiplied by 10, and the phase of this function is a constant ; N(90°). The magnitude and phase characteristics for poles and zeros at the origin are shown in Figs. 12.10b and c, respectively. Simple Pole or Zero Linear approximations can be employed when a simple pole or zero of the form (1 + j) is present in the network function. For V 1, (1 + j) L 1, and therefore, 20 log 10∑(1 + j)∑ = 20 log 10 1 = 0 dB. Similarly, if W 1, then (1 + j) L j, and hence 20 log 10∑(1 + j)∑ L 20 log 10 . Therefore, for V 1 the response is 0 dB and for W 1 the response has a slope that is the same as that of a simple pole or zero at the origin. The intersection of these two asymptotes, one for V 1 and one for W 1, is the point where = 1 or = 1兾, which is called the break frequency. At this break frequency, where = 1兾, 20 log 10 @(1 + j1)@ = 20 log 10(2)1兾2 = 3 dB. Therefore, the actual curve deviates from the asymptotes by 3 dB at the break frequency. It can be shown that at 6-08-2010 15:14 Page 589 SECTION 12.2 Magnitude (dB) 1 2 3 4 5 6 7 8 91 3 4 5 6 7 8910 –45°/decade –40 –45° –60 –12 –80 –90° –18 –20 0.2 0.5 1.0 2.0 10 4.0 (rad/s)(Log scale) (a) 90° =tan–1 ±12 60 45° ±6 dB ±45°/decade ±20 dB/decade 30 Phase shift (deg) Magnitude (dB) ±18 dB=20 log10|(1+j)| 0 0 0.1 0.2 0.5 1.0 2.0 4.0 589 Figure 12.11 dB=20 log10|(1+j)–1| 0 = tan–1 dB –20 dB/decade –20 0 –6 2 S I N U S O I D A L F R E Q U E N C Y A N A LY S I S Phase shift (deg) irwin12_577-666hr.qxd 10 (rad/s)(Log scale) (b) one-half and twice the break frequency, the deviations are 1 dB. The phase angle associated with a simple pole or zero is = tan-1 , which is a simple arctangent curve. Therefore, the phase shift is 45° at the break frequency and 26.6° and 63.4° at one-half and twice the break frequency, respectively. The actual magnitude curve for a pole of this form is shown in Fig. 12.11a. For a zero the magnitude curve and the asymptote for W 1 have a positive slope, and the phase curve extends from 0° to +90°, as shown in Fig. 12.11b. If multiple poles or zeros of the form (1 + j)N are present, then the slope of the high-frequency asymptote is multiplied by N, the deviation between the actual curve and the asymptote at the break frequency is 3N dB, and the phase curve extends from 0 to N(90°) and is N(45°) at the break frequency. Quadratic Poles or Zeros Quadratic poles or zeros are of the form 1 + 2 (j) + (j)2 This term is a function not only of but also of the dimensionless term , which is called the damping ratio. If >1 or = 1, the roots are real and unequal or real and equal, respectively, and these two cases have already been addressed. If <1, the roots are complex conjugates, and it is this case that we will examine now. Following the preceding argument for a simple pole or zero, the log magnitude of the quadratic factor is 0 dB for V 1. For W 1, 20 log 10 @ 1 - ()2 + 2j() @ L 20 log 10 @()2 @ = 40 log 10∑∑ and therefore, for W 1, the slope of the log magnitude curve is + 40 db/decade for a quadratic zero and -40 db/decade for a quadratic pole. Between the two extremes, V 1 Magnitude and phase plot (a) for a simple pole, and (b) for a simple zero. irwin12_577-666hr.qxd 590 6-08-2010 CHAPTER 12 15:14 Page 590 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E 1 Figure 12.12 3 4 5 6 7 8 91 2 3 4 5 6 7 8 910 =0 20log10|[1±2(j)+(j)2]–1| =0.1 20 Magnitude (dB) Magnitude and phase characteristics for quadratic poles. 2 0.2 10 0.6 0.8 1.0 0.4 0 –10 –20 –30 0.2 0.5 1.0 2.0 4.0 10 (rad/s)(Log scale) (a) 1 2 3 4 5 6 7891 =0.1 Phase shift (deg) 0 –40 2 3 4 5 6 7 8 9 10 0.2 2 0.4 =–tan–1 — 1–()2 0.6 0.8 1.0 –80 –120 –160 –200 0.2 0.5 1.0 2.0 4.0 10 (rad/s)(Log scale) (b) and W 1, the behavior of the function is dependent on the damping ratio . Fig. 12.12a illustrates the manner in which the log magnitude curve for a quadratic pole changes as a function of the damping ratio. The phase shift for the quadratic factor is tan-1 2兾 C1 - ()2 D. The phase plot for quadratic poles is shown in Fig. 12.12b. Note that in this case the phase changes from 0° at frequencies for which V 1 to –180° at frequencies for which W 1. For quadratic zeros the magnitude and phase curves are inverted; that is, the log magnitude curve has a slope of +40 db/decade for W 1, and the phase curve is 0° for V 1 and +180° for W 1. EXAMPLE 12.3 We want to generate the magnitude and phase plots for the transfer function Gv(j) = SOLUTION 10(0.1j + 1) (j + 1)(0.02j + 1) Note that this function is in standard form, since every term is of the form (j + 1). To determine the composite magnitude and phase characteristics, we will plot the individual asymptotic terms and then add them as specified in Eqs. (12.8) and (12.9). Let us consider the magnitude plot first. Since K0=10, 20 log 10 10 = 20 dB, which is a constant independent of irwin12_577-666hr.qxd 6-08-2010 15:14 Page 591 SECTION 12.2 S I N U S O I D A L F R E Q U E N C Y A N A LY S I S 591 frequency, as shown in Fig. 12.13a. The zero of the transfer function contributes a term of the form +20 log 10∑1 + 0.1j∑, which is 0 dB for 0.1 V 1, has a slope of ±20 dB/decade for 0.1 W 1, and has a break frequency at =10 rad/s. The poles have break frequencies at =1 and =50 rad/s. The pole with break frequency at =1 rad/s contributes a term of the form -20 log 10∑1 + j∑, which is 0 dB for V 1 and has a slope of –20 dB/decade for W 1. A similar argument can be made for the pole that has a break frequency at = 50 rad兾s. These factors are all plotted individually in Fig. 12.13a. Consider now the individual phase curves. The term K0 is not a function of and does not contribute to the phase of the transfer function. The phase curve for the zero is +tan-1 0.1, which is an arctangent curve that extends from 0° for 0.1 V 1 to ±90° for 0.1 W 1 and has a phase of ±45° at the break frequency. The phase curves for the two poles are -tan-1 and -tan-1 0.02 . The term -tan-1 is 0° for V 1, –90° for W 1, and –45° at the break frequency =1 rad/s. The phase curve for the remaining pole is plotted in a similar fashion. All the individual phase curves are shown in Fig. 12.13a. 1 3 5 7 91 3 5 7 91 3 5 7 91 3 5 7 91 Figure 12.13 40 20 log10(10) 20 –20 log10|1±0.02j| –20 log10|1±j| ±90° –20 ±45° ±tan–10.1 0° –tan–1 –tan–10.02 Phase shift (deg) Log magnitude (dB) 20 log10|1±0.1j| 0 –45° –90° 0.1 1.0 10 100 1000 (rad/s) (a) 1 3 5 7 91 3 5 7 91 3 5 7 91 3 5 7 91 40 Composite magnitude 0 ±90° –20 ±45° 0° Composite phase 0.1 1.0 10 (rad/s) (b) 100 –45° –90° 1000 Phase shift (deg) Log magnitude (dB) 20 (a) Magnitude and phase components for the poles and zeros of the transfer function in Example 12.3; (b) Bode plot for the transfer function in Example 12.3. irwin12_577-666hr.qxd 592 6-08-2010 CHAPTER 12 15:14 Page 592 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E As specified in Eqs. (12.8) and (12.9), the composite magnitude and phase of the transfer function are obtained simply by adding the individual terms. The composite curves are plotted in Fig. 12.13b. Note that the actual magnitude curve (solid line) differs from the straight-line approximation (dashed line) by 3 dB at the break frequencies and 1 dB at one-half and twice the break frequencies. Let us draw the Bode plot for the following transfer function: 12.4 Gv(j) = Once again all the individual terms for both magnitude and phase are plotted in Fig. 12.14a. The straight line with a slope of –40 dB/decade is generated by the double pole at the Figure 12.14 1 3 5 7 91 3 5 7 91 3 5 7 91 40 20 log10(25)=28 dB –40 log10|j| Log magnitude (dB) 20 ±20 log10|1+j| –20 log10|1+0.1j| 0 –20 ±90° ±tan–1 0° –tan–10.1 –90° –180° 0.1 1.0 –270° 100 10 (rad/s) (a) 1 3 5 7 91 3 5 7 91 3 5 7 91 60 40 Log magnitude (dB) (a) Magnitude and phase components for the poles and zeros of the transfer function in Example 12.4; (b) Bode plot for the transfer function in Example 12.4. (j)2(0.1j + 1) Phase shift (deg) SOLUTION 25(j + 1) Composite magnitude 20 0 0° Composite phase –90° Phase shift (deg) EXAMPLE –180° 0.1 1.0 10 (rad/s) (b) –270° 100 irwin12_577-666hr.qxd 6-08-2010 15:14 Page 593 SECTION 12.2 S I N U S O I D A L F R E Q U E N C Y A N A LY S I S 593 origin. This line is a plot of –40 log10 versus and therefore passes through 0 dB at =1 rad/s. The phase for the double pole is a constant –180° for all frequencies. The remainder of the terms are plotted as illustrated in Example 12.3. The composite plots are shown in Fig. 12.14b. Once again they are obtained simply by adding the individual terms in Fig. 12.14a. Note that for frequencies for which V 1, the slope of the magnitude curve is –40 dB/decade. At =1 rad/s, which is the break frequency of the zero, the magnitude curve changes slope to –20 dB/decade. At =10 rad/s, which is the break frequency of the pole, the slope of the magnitude curve changes back to –40 dB/decade. The composite phase curve starts at –180° due to the double pole at the origin. Since the first break frequency encountered is a zero, the phase curve shifts toward –90°. However, before the composite phase reaches –90°, the pole with break frequency =10 rad/s begins to shift the composite curve back toward –180°. Example 12.4 illustrates the manner in which to plot directly terms of the form K0兾(j)N. For terms of this form, the initial slope of –20N dB/decade will intersect the 0-dB axis at a frequency of AK0 B 1兾N rad兾s; that is, -20 log 10 @K0兾(j)N @ = 0 dB implies that K0兾(j)N = 1, and therefore, = AK0 B 1兾N rad兾s. Note that the projected slope of the magnitude curve in Example 12.4 intersects the 0-dB axis at = (25)1兾2 = 5 rad兾s. Similarly, it can be shown that for terms of the form K0(j)N, the initial slope 1兾N of ±20N dB/decade will intersect the 0-dB axis at a frequency of = A1兾K0 B rad兾s; N N that is, +20 log 10 @K0兾(j) @ = 0 dB implies that K0兾(j) = 1, and therefore = 1兾N A1兾K0 B rad兾s. By applying the concepts we have just demonstrated, we can normally plot the log magnitude characteristic of a transfer function directly in one step. Learning Assessments E12.4 Sketch the magnitude characteristic of the Bode plot, labeling all critical slopes and points for the function G(j) = 104(j + 2) ANSWER: |G| (dB) ±20 dB/decade (j + 10)(j + 100) –20 dB/decade 20log1020 2 Figure E12.4 E12.5 Sketch the magnitude characteristic of the Bode plot for the 10 100 (rad/s) ANSWER: transfer function: H(j) = |H| (dB) 5(j + 10) j(j + 100) –6 –20 dB/decade –20 dB/decade –26 Figure E12.5 1 10 100 (rad/s) irwin12_577-666hr.qxd 594 6-08-2010 CHAPTER 12 15:14 Page 594 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E E12.6 Sketch the magnitude characteristic of the Bode plot, labeling ANSWER: all critical slopes and points for the function G(j) = |G| (dB) 100(0.02j + 1) –40 dB/decade (j)2 0 –20 dB/decade 10 Figure E12.6 50 (rad/s) E12.7 Sketch the magnitude characteristic of the Bode plot, labeling ANSWER: all critical slopes and points for the function G(j) = |G| (dB) 10j 0 (j + 1)(j + 10) –20 dB/decade ±20 dB/decade Figure E12.7 EXAMPLE 10 (rad/s) We wish to generate the Bode plot for the following transfer function: 12.5 Gv(j) = SOLUTION 1 25j (j + 0.5) C (j)2 + 4j + 100 D Expressing this function in standard form, we obtain Gv(j) = 0.5j (2j + 1) C(j兾10)2 + j兾25 + 1 D The Bode plot is shown in Fig. 12.15. The initial low-frequency slope due to the zero at the origin is ±20 dB/decade, and this slope intersects the 0-dB line at =1/K0=2 rad/s. At =0.5 rad/s the slope changes from ±20 dB/decade to 0 dB/decade due to the presence of the pole with a break frequency at =0.5 rad/s. The quadratic term has a center frequency of =10 rad/s (i.e., =1/10). Since 2 = 1 25 and = 0.1 then = 0.2 irwin12_577-666hr.qxd 6-08-2010 15:14 Page 595 SECTION 12.2 1 3 5 7 91 3 5 7 91 3 5 7 91 S I N U S O I D A L F R E Q U E N C Y A N A LY S I S 3 5 7 91 Figure 12.15 Bode plot for the transfer function in Example 12.5. 20 0 –20 –40 ±90 0 Phase shift (deg) Log magnitude (dB) 595 –90 –180 0.01 0.1 1.0 (rad/s) 10 100 Plotting the curve in Fig. 12.12a with a damping ratio of =0.2 at the center frequency =10 rad/s completes the composite magnitude curve for the transfer function. The initial low-frequency phase curve is ±90°, due to the zero at the origin. This curve and the phase curve for the simple pole and the phase curve for the quadratic term, as defined in Fig. 12.12b, are combined to yield the composite phase curve. Learning Assessment E12.8 Given the following function G(j), sketch the magnitude characteristic of the Bode plot, labeling all critical slopes and points: ANSWER: |G| (dB) G(j) = 0.2(j + 1) j C(j兾12)2 + j兾36 + 1D 0 1 =— 6 –20 dB/decade –40 dB/decade Figure E12.8 D E R I V I N G T H E T R A N S F E R F U N C T I O N F R O M T H E B O D E P LOT 0.2 1 12 (rad/s) The following example will serve to demonstrate the derivation process. Given the asymptotic magnitude characteristic shown in Fig. 12.16, we wish to determine the transfer function Gv(j). Since the initial slope is 0 dB/decade, and the level of the characteristic is 20 dB, the factor K0 can be obtained from the expression EXAMPLE SOLUTION 12.6 irwin12_577-666hr.qxd 596 6-08-2010 15:14 CHAPTER 12 Page 596 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E 20 dB=20 log10 K0 and hence K0=10 1 Figure 12.16 Straight-line magnitude plot employed in Example 12.6. 3 5 7 91 3 5 7 91 3 5 7 91 3 5 7 91 20 –20dB/decade Magnitude (dB) 0 –20dB/decade –20 –40dB/decade 0.01 0.1 1.0 (rad/s) 10.0 100.0 The –20-dB/decade slope starting at =0.1 rad/s indicates that the first pole has a break frequency at =0.1 rad/s, and therefore one of the factors in the denominator is (10j + 1). The slope changes by ±20 dB/decade at =0.5 rad/s, indicating that there is a zero present with a break frequency at =0.5 rad/s, and therefore the numerator has a factor of (2j + 1). Two additional poles are present with break frequencies at =2 rad/s and =20 rad/s. Therefore, the composite transfer function is Gv(j) = 10(2j + 1) (10j + 1)(0.5j + 1)(0.05j + 1) Note carefully the ramifications of this example with regard to network design. Learning Assessments E12.9 Determine the transfer function G(j) if the straight-line magnitude ANSWER: characteristic approximation for this function is as shown in Fig. E12.9. |G| (dB) G(j) = –20 dB/decade –20 dB/decade 0 dB –20 dB/decade Figure E12.9 5 20 50 100 (rad/s) 5a j j + 1b a + 1b 5 50 j j j a + 1b a + 1b 20 100 . irwin12_577-666hr.qxd 6-08-2010 15:14 Page 597 SECTION 12.3 H(j) = 20 dB/dec 597 ANSWER: E12.10 Find H(j) if its magnitude characteristic is shown in Fig. E12.10. |H|(dB) RESONANT CIRCUITS 0.3162 a j + 1b 5 j j j + 1b a + 1b a + 1b a 12 25 40 . –20 dB/dec –10 –40 dB/dec Figure E12.10 5 12 25 40 (rad/s) 12.3 SERIES RESONANCE A circuit with extremely important frequency characteristics is shown in Fig. 12.17. The input impedance for the series RLC circuit is Z(j) = R + jL + 1 1 b = R + j a L jC C 12.10 Resonant Circuits The imaginary term will be zero if L = 1 C I The value of that satisfies this equation is 1 0 = 1LC + VR 12.11 R - and at this value of the impedance becomes ZAj0 B = R + V1 C 12.12 This frequency 0 , at which the impedance of the circuit is purely real, is also called the resonant frequency, and the circuit itself at this frequency is said to be in resonance. Resonance is a very important consideration in engineering design. For example, engineers designing the attitude control system for the Saturn vehicles had to ensure that the control system frequency did not excite the body bending (resonant) frequencies of the vehicle. Excitation of the bending frequencies would cause oscillations that, if continued unchecked, would result in a buildup of stress until the vehicle would finally break apart. Resonance is also a benefit, providing string and wind musical instruments with volume and rich tones. At resonance the voltage and current are in phase and, therefore, the phase angle is zero and the power factor is unity. At resonance the impedance is a minimum and, therefore, the current is maximum for a given voltage. Fig. 12.18 illustrates the frequency response of the series RLC circuit. Note that at low frequencies the impedance of the series circuit is dominated by the capacitive term, and at high frequencies the impedance is dominated by the inductive term. Resonance can be viewed from another perspective—that of the phasor diagram. In the series circuit the current is common to every element. Therefore, the current is employed as reference. The phasor diagram is shown in Fig. 12.19 for the three frequency values 6 0 , = 0 , 7 0 . L - Figure 12.17 Series RLC circuit. irwin12_577-666hr.qxd 598 6-08-2010 CHAPTER 12 15:14 Page 598 VA R I A B L E - F R E Q U E N C Y N E T W O R K P E R F O R M A N C E When 6 0 , VC 7 VL , Z is negative and the voltage V1 lags the current. If = 0 , VL = VC , Z is zero, and the voltage V1 is in phase with the current. If 7 0 , VL 7 VC , Z is positive, and the voltage V1 leads the current. For the series circuit we define what is commonly called the quality factor Q as Q = 0 L 1 1 L = = R 0 CR R BC 12.13 Q is a very important factor in resonant circuits, and its ramifications will be illustrated throughout the remainder of this section. Figure 12.18 |Z| Frequency response of a series RLC circuit. |Z| R 0 0 1 L-—— C Figure 12.19 Phasor diagrams for the series RLC circuit. VL VL VL VR V1 VR=V1 I I Z Z VC EXAMPLE 12.7 SOLUTION I VR V1 VC VC <0 =0 >0 Consider the network shown in Fig. 12.20. Let us determine the resonant frequency, the voltage across each element at resonance, and the value of the quality factor. The resonant frequency is obtained from the expression 0 = = 1 1LC 1 2(25)A10-3 B(10)A10-6 B = 2000 rad兾s At this resonant frequency I = V V = = 5 / 0° A Z R irwin12_577-666hr.qxd 6-08-2010 15:14 Page 599 SECTION 12.3 Therefore, RESONANT CIRCUITS 599 VR = A5 / 0°B(2) = 10 / 0° V VL = j0 LI = 250 / 90° V VC = I = 250 / -90° V j0 C Note the magnitude of the voltages across the inductor and capacitor with respect to the input voltage. Note also that these voltages are equal and are 180° out of phase with one another. Therefore, the phasor diagram for